Nutrient Requirement


[PDF]Nutrient Requirement - Rackcdn.comc813587.r87.cf2.rackcdn.com/...

0 downloads 165 Views 2MB Size

http://www.nap.edu/catalog/6016.html We ship printed books within 1 business day; personal PDFs are available immediately.

Nutrient Requirements of Swine: 10th Revised Edition Subcommittee on Swine Nutrition, Committee on Animal Nutrition, National Research Council ISBN: 0-309-54988-4, 210 pages, 8.5 x 11, (1998) This PDF is available from the National Academies Press at: http://www.nap.edu/catalog/6016.html

Visit the National Academies Press online, the authoritative source for all books from the National Academy of Sciences, the National Academy of Engineering, the Institute of Medicine, and the National Research Council: • Download hundreds of free books in PDF • Read thousands of books online for free • Explore our innovative research tools – try the “Research Dashboard” now! • Sign up to be notified when new books are published • Purchase printed books and selected PDF files

Thank you for downloading this PDF. If you have comments, questions or just want more information about the books published by the National Academies Press, you may contact our customer service department tollfree at 888-624-8373, visit us online, or send an email to [email protected].

This book plus thousands more are available at http://www.nap.edu. Copyright © National Academy of Sciences. All rights reserved. Unless otherwise indicated, all materials in this PDF File are copyrighted by the National Academy of Sciences. Distribution, posting, or copying is strictly prohibited without written permission of the National Academies Press. Request reprint permission for this book.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

NATIONAL RESEARCH COUNCIL

NUTRIENT REQUIREMENTS OF SWINE Tenth Revised Edition 1998

NUTRIENT REQUIREMENTS OF DOMESTIC ANIMALS

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirements of Swine Tenth Revised Edition, 1998

Subcommittee on Swine Nutrition Committee on Animal Nutrition Board on Agriculture National Research Council

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

NATIONAL ACADEMY PRESS ● 2101 Constitution Avenue, NW ● Washington, D.C. 20418 NOTICE: The project that is the subject of this report was approved by the Governing Board of the National Research Council, whose members are drawn from the councils of the National Academy of Sciences, the National Academy of Engineering, and the Institute of Medicine. The members of the committee responsible for the report were chosen for their special competencies and with regard for appropriate balance. This study was supported by the Agricultural Research Service of the U.S. Department of Agriculture, under Agreement No. 59-32U4-5-6, and by the Center for Veterinary Medicine, Food and Drug Administration of the U.S. Department of Health and Human Services, under Cooperative Agreement No. FD-U-000006-10. Additional support was provided by the American Feed Industry Association, and the National Pork Producers Council. The National Academy of Sciences is a private, nonprofit, self-perpetuating society of distinguished scholars engaged in scientific and engineering research, dedicated to the furtherance of science and technology and to their use for the general welfare. Upon the authority of the charter granted to it by the Congress in 1863, the Academy has a mandate that requires it to advise the federal government on scientific and technical matters. Dr. Bruce Alberts is president of the National Academy of Sciences. The National Academy of Engineering was established in 1964, under the charter of the National Academy of Sciences, as a parallel organization of outstanding engineers. It is autonomous in its administration and in the selection of its members, sharing with the National Academy of Sciences the responsibility for advising the federal government. The National Academy of Engineering also sponsors engineering programs aimed at meeting national needs, encourages education and research, and recognizes the superior achievements of engineers. Dr. William A. Wulf is acting president of the National Academy of Engineering. The Institute of Medicine was established in 1970 by the National Academy of Sciences to secure the services of eminent members of appropriate professions in the examination of policy matters pertaining to the health of the public. The Institute acts under the responsibility given to the National Academy of Sciences by its congressional charter to be an adviser to the federal government and, upon its own initiative, to identify issues of medical care, research, and education. Dr. Kenneth I. Shine is president of the Institute of Medicine. The National Research Council was organized by the National Academy of Sciences in 1916 to associate the broad community of science and technology with the Academy’s purposes of furthering knowledge and advising the federal government. Functioning in accordance with general policies determined by the Academy, the Council has become the principal operating agency of both the National Academy of Sciences and the National Academy of Engineering in providing services to the government, the public, and the scientific and engineering communities. The Council is administered jointly by both Academies and the Institute of Medicine. Dr. Bruce Alberts and Dr. William A. Wulf are chairman and vice-chairman, respectively, of the National Research Council. Library of Congress Cataloging-in-Publication Data Nutrient requirements of swine / Subcommittee on Swine Nutrition, Committee on Animal Nutrition, Board on Agriculture, National Research Council. — 10th rev. ed. p. cm. — (Nutrient requirements of domestic animals) Includes bibliographical references and index. ISBN 0-309-05993-3 (pbk.) 1. Swine—Nutrition—Requirements. 2. Swine—Feeding and feeds. I. National Research Council (U.S.). Subcommittee on Swine Nutrition. II. Series: Nutrient requirements of domestic animals (Unnumbered) SF396.5 .N87 1988 636.480852—ddc21 98-9007 CIP International Standard Book Number 0-309-05993-3 ©1998 by the National Academy of Sciences. All rights reserved. No part of this book may be reproduced by any mechanical, photographic, or electronic process, or in the form of a phonographic recording, nor may it be stored in a retrieval system, transmitted, or otherwise copied for public or private use without written permission from the publisher, except for the purposes of official use by the U.S. government. Additional copies of this report are available from National Academy Press, 2101 Constitution Avenue, N.W., Lockbox 285, Washington, D.C. 20055; (800) 624-6242 or (202) 334-3313 (in the Washington metropolitan area); Internet, http://www.nap.edu Printed in the United States of America. This report and the computer model are also available on the Internet, http://www.nap.edu/readingroom/ enter2.cgi?0309059933.html.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

SUBCOMMITTEE ON SWINE NUTRITION GARY L. CROMWELL, Chair, University of Kentucky DAVID H. BAKER, University of Illinois RICHARD C. EWAN, Iowa State University E.T. KORNEGAY, Virginia Polytechnic Institute and State University AUSTIN J. LEWIS, University of Nebraska JAMES E. PETTIGREW, Pettigrew Consulting International, Louisiana, Missouri NORMAN C. STEELE, U.S. Department of Agriculture, Agricultural Research Service, Beltsville, Maryland PHILIP A. THACKER, University of Saskatchewan, Canada COMMITTEE ON ANIMAL NUTRITION DONALD C. BEITZ, Chairman, Iowa State University GARY L. CROMWELL, University of Kentucky* GEORGE C. FAHEY, University of Illinois*** DELBERT M. GATLIN III, Texas A&M University RONALD L. HORST, U.S. Department of Agriculture, Agricultural Research Service, Ames, Iowa*** TERRY J. KLOPFENSTEIN, University of Nebraska*** LAURIE M. LAWRENCE, University of Kentucky* AUSTIN J. LEWIS, University of Nebraska CARL M. PARSONS, University of Illinois ALICE N. PELL, Cornell University*** GARY D. POTTER, Texas A&M University JERRY L. SELL, Iowa State University** ROBERT P. WILSON, Mississippi State University** KARIN M. WITTENBERG, University of Manitoba, Canada *July 1, 1992, through June 30, 1995 **July 1, 1993, through June 30, 1996 ***July 1, 1994, through June 30, 1997

Staff CHARLOTTE KIRK BAER, Program Director MELINDA SIMONS, Project Assistant

iii

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

BOARD ON AGRICULTURE DALE E. BAUMAN, Chair, Cornell University JOHN M. ANTLE, Montana State University SANDRA S. BATIE, Michigan State University MAY R. BERENBAUM, University of Illinois LEONARD S. BULL, North Carolina State University WILLIAM B. DELAUDER, Delaware State College ANTHONY S. EARL, Quarles & Brady Law Firm, Madison, Wisconsin ESSEX E. FINNEY, JR., U.S. Department of Agriculture (retired), Mitchellville, Maryland CORNELIA B. FLORA, Iowa State University GEORGE R. HALLBERG, University of Iowa RICHARD R. HARWOOD, Michigan State University T. KENT KIRK, University of Wisconsin, Madison HARLEY W. MOON, Iowa State University WILLIAM L. OGREN, University of Illinois GEORGE E. SEIDEL, JR., Colorado State University JOHN W. SUTTIE, University of Wisconsin JAMES J. ZUICHES, Washington State University J. PAUL GILMAN, Executive Director MICHAEL J. PHILLIPS, Director

v

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Preface The model for growing-finishing pigs allows the user to generate tables of nutrient requirements for various body weights of pigs, based on the pig’s lean growth rate, gender, and environmental conditions. Similarly, the energy and amino acid requirements of gestating and lactating sows are estimated by models, and the user can generate nutrient requirement tables for sows with different body weights and weight gains during gestation and for various levels of lactational productivity. To accomplish this, a user-friendly computer program containing the models is included in this edition. Requirements for amino acids in the models were generated on a true ileal digestible basis. The amino acid requirements are provided to the user on a true and apparent digestible basis as well as on a total basis, using corn and soybean meal as the major ingredients. The models also estimate energy requirements for gestating and lactating sows and energy intakes of growing pigs given ad libitum access to feed. Equations to estimate mineral and vitamin requirements at various body weights are also included in the growth model. Other new information is presented in this tenth edition. Minimizing nutrient excretion is addressed and a discussion of nonnutritive feed additives was expanded. New information on the nutrient composition of an expanded list of feed ingredients and on the bioavailability of amino acids (true and apparent ileal basis), phosphorus, and other nutrients is also included in this edition. Finally, the nutrient requirement tables also provide more information than did those in previous editions. This three-year study was conducted by the Subcommittee on Swine Nutrition, which was appointed in 1994 under the guidance of the Board on Agriculture’s Committee on Animal Nutrition. The subcommittee began its work in November 1994 and the study was completed in December 1997, with the release of the report in April 1998.

Swine production represents an important segment of the food animal industry in the United States and throughout the world. Pork is an important source of energy, protein, minerals, and vitamins, and is the most widely consumed red meat in the world. Proper formulation of diets is fundamental to the efficient production of swine in systems that address environmental concerns, and this process depends on a knowledge of the nutrient requirements of swine and the nutritional characteristics of nutrient sources. This tenth edition of Nutrient Requirements of Swine contains a reassessment of the nutrient requirements of swine and incorporates new information that was used to establish the requirements. An abundance of new knowledge in swine nutrition has surfaced since the last edition of Nutrient Requirements of Swine was published in 1988. There is now a greater awareness and understanding of the effects of growth rate, carcass leanness, gender, health, environmental temperature, crowding, and carcass modifiers on the nutrient requirements of growing pigs. The higher nutrient requirements of prolific sows nursing large litters are now better understood. Additionally, new information on the bioavailability of nutrients is now available. A better understanding of the nutrient requirements and nutrient sources allows one to accurately formulate diets to meet the pig’s dietary requirements without producing overages of nutrients that are excreted into the environment. A major change was made in this edition in that the subcommittee provided the biological basis used to establish energy and amino acid requirements in the form of integrated mathematical equations (models). The models were developed by the subcommittee with the goal of keeping them simple, transparent (i.e., inner parts understandable to the user), and firmly anchored to empirical data. The process of model development and validation was an extremely laborious and time-consuming task. While these versions of the models are not perfect, the subcommittee believes that they represent a marked improvement over previous systems of establishing requirements and provide the groundwork for development of improved models by future subcommittees.

GARY L. CROMWELL, Chair Subcommittee on Swine Nutrition

vii

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Acknowledgments

The subcommittee would like to acknowledge the many scientists who conducted the research studies from which we were able to draw information to establish nutrient requirements. Appreciation is given to Kevin Halpin, chairman of the Nutrition Council Swine Committee, American Feed Industry Association, who assisted our subcommittee in setting goals and establishing direction for the revised publication, and to the many companies and individuals in the feed industry who provided useful information to the subcommittee. The subcommittee thanks Tim Stahly, Iowa State University, and Allan Schinckel, Purdue University, for insights gained from discussions with them during the initial stages of model development. The subcommittee wishes to thank Charlotte Kirk Baer, Program Director, Board on Agriculture, for her untiring efforts in seeing this project to completion. Appreciation is also given to staff members Melinda Simons and Juliemarie Goupil for their assistance with the report, and to Mary Poos for her help during the first year of planning. Finally, the work by Ron Haugen, Easy Systems, Inc., in developing the software interface for the model is acknowledged. The generous support of this study provided by the National Pork Producers Council is gratefully acknowledged. In addition, the subcommittee appreciates the support provided by the U.S. Department of Agriculture’s Agricultural Research Service, the Center for Veterinary

Medicine of the Department of Health and Human Service’s Food and Drug Administration, and the American Feed Industry Association. This report has been reviewed by individuals chosen for their diverse perspectives and technical expertise, in accordance with procedures approved by the National Research Council’s (NRC) Report Review Committee. The purpose of this independent review is to provide candid and critical comments that will assist the authors and the NRC in making the published report as sound as possible and to ensure that the report meets institutional standards for objectivity, evidence, and responsiveness to the study charge. The content of the review comments and draft manuscript remain confidential to protect the integrity of the deliberative process. We wish to thank the following individuals for their participation in the review of this report: R. Dean Boyd, Pig Improvement Company, USA; Thomas Crenshaw, University of Wisconsin; C.F.M. deLange, University of Guelph; Darrell Knabe, Texas A&M University; Harley W. Moon, Iowa State University; Robert Myer, University of Florida; Carl Parsons, University of Illinois; Tim Stahly, Iowa State University; Michael Tokach, Kansas State University; and Gawain Willis, Purina Mills, Inc. While the individuals listed above have provided many constructive comments and suggestions, responsibility for the final content of this report rests solely with the authoring committee and the NRC.

ix

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Contents PREFACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .vii ACKNOWLEDGMENTS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix OVERVIEW . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1 1 ENERGY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3 Classification of Energy, 3 Gross Energy, 3 Digestible Energy, 3 Metabolizable Energy, 4 Net Energy, 4 Heat Production, 5 Temperature, 5 Activity, 5 Energy Requirements, 5 Maintenance, 5 Growth, 6 Pregnancy, 6 Lactation, 7 Developing Boars and Gilts, 7 Sexually Active Boars, 8 Energy Sources, 8 Sugars and Starch, 8 Nonstarch Polysaccharides, 8 Lipids, 9 Voluntary Feed Intake, 10 Suckling Pigs, 10 Weanling Pigs, 10 Growing-Finishing Pigs, 10 Sows, 11 References, 11

2 PROTEINS AND AMINO ACIDS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .16 Essential and Nonessential Amino Acids, 16

xi

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

xii

Contents Amino Acids in Diets, 17 Ratios Among Amino Acids (Ideal Protein), 17 Bioavailability of Amino Acids, 18 Amino Acid Isomers, 19 Amino Acid Deficiencies and Excesses, 19 Amino Acid Requirements, 19 Starting Pigs, 19 Growing-Finishing Pigs, 24 Sows, 25 Boars, 25 References, 26

3 MODELS FOR ESTIMATING ENERGY AND AMINO ACID REQUIREMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .31 Overview of the Models, 31 Growing-Finishing Pigs, 32 Requirement for Lysine, 32 Requirements for Other Amino Acids, 35 Expression of Amino Acid Requirements, 35 Gestating Sows, 36 Composition of Weight Gain, 37 Requirement for Energy, 37 Requirement for Lysine, 38 Requirements for Other Amino Acids, 38 Expression of Amino Acid Requirements, 38 Lactating Sows, 38 Requirement for Energy, 39 Requirement for Lysine, 39 Requirements for Other Amino Acids, 40 Expression of Amino Acid Requirements, 40 Weanling Pigs, 40 Mineral and Vitamin Requirements, 41 Evaluation of the Models, 42 Growth Model, 42 Gestation Model, 44 Lactation Model, 44 References, 44

4 MINERALS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .47 Macro Minerals, 47 Calcium and Phosphorus, 47 Sodium and Chlorine, 49 Magnesium, 50 Potassium, 51 Sulfur, 51 Micro/Trace Minerals, 51 Chromium, 51 Cobalt, 52 Copper, 52 Iodine, 53 Iron, 54 Manganese, 55

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Contents

xiii

Selenium, 55 Zinc, 56 References, 57

5 VITAMINS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .71 Fat-Soluble Vitamins, 71 Vitamin A, 71 Vitamin D, 73 Vitamin E, 73 Vitamin K, 74 Water-Soluble Vitamins, 75 Biotin, 75 Choline, 76 Folacin, 77 Niacin, 78 Pantothenic Acid, 78 Riboflavin, 79 Thiamin, 79 Vitamin B6 (The Pyridoxines), 80 Vitamin B12, 80 Vitamin C (Ascorbic Acid), 81 References, 82

6 WATER . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .90 Functions of Water, 90 Water Turnover, 90 Water Requirements, 91 Suckling Pigs, 91 Weanling Pigs, 91 Growing-Finishing Pigs, 92 Gestating Sows, 93 Lactating Sows, 93 Boars, 93 Water Quality, 93 References, 95

7 NONNUTRITIVE FEED ADDITIVES . . . . . . . . . . . . . . . . . . . . . . . . . . . .97 Additives, 97 Antimicrobial Agents, 97 Anthelmintics, 98 Microbial Supplements, 98 Oligosaccharides, 98 Enzymes, 98 Acidifiers, 98 Flavors, 99 Odor Control Agents, 99 Antioxidants, 99 Pellet Binders, 99 Flow Agents, 99 Mineral Supplements, 99 Carcass Modifiers, 99 Safety Concerns, 99

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

xiv

Contents Regulations, 100 References, 100

8 MINIMIZING NUTRIENT EXCRETION . . . . . . . . . . . . . . . . . . . . . . . . .103 References, 105

9 DIET FORMULATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .107 Formulating a Corn–Soybean Meal Diet, 107 Formulation, 107

10 NUTRIENT REQUIREMENT TABLES . . . . . . . . . . . . . . . . . . . . . . . . .110 11 COMPOSITION OF FEED INGREDIENTS . . . . . . . . . . . . . . . . . . . . .124 References, 142

APPENDIXES 1 Equations Used to Model the Biological Basis for Predicting Nutrient Requirements, 143 2 Equations for Determining Lean Growth Rate of Pigs, 148 3 Method to Create a Cubic Regression Equation, 150 4 A User’s Guide for Model Application, 153 5 Help Screens, 171

AUTHORS. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .179 INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .180

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Tables and Figures

TABLES 2-1

Ideal Ratios of Amino Acids to Lysine for Maintenance, Protein Accretion, Milk Synthesis, and Body Tissue, 18

2-2

Research Findings on Amino Acid Requirements of Growing Swine Since 1985, 20

2-3

Lysine Requirements of Gestating and Lactating Sows, 25

3-1

Equations for Converting Percentages of Amino Acids from a True Ileal Digestible Basis to an Apparent Ileal Digestible Basis, from an Apparent Ileal Digestible Basis to a True Ileal Digestible Basis, and from a True or Apparent Ileal Digestible Basis to a Total Basis in a Corn–Soybean Meal Diet, 36

3-2

Coefficients Used in the Growth Model to Predict Mineral and Vitamin Requirements (percentage or amount/kg of diet) for Pigs of Various Body Weights, 43

3-3

Evaluation of Data of the Growth Model, 43

3-4

Evaluation of Data of the Lactation Model, 44

6-1

Evaluation of Water Quality for Pigs Based on Total Dissolved Solids, 94

6-2

Water Quality Guidelines for Livestock, 95

9-1

Nutrients in Corn and Corn ` Soybean Meal (Dehulled) Compared with the Nutrient Requirements of a 40-kg Growing Pig of High-Medium Lean Growth Rate (325 g of carcass fat-free lean/day), 108

9-2

Fortified Swine Diet, 109

10-1

Dietary Amino Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90 percent dry matter), 111

10-2

Daily Amino Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90 percent dry matter), 112

10-3

Dietary Amino Acid Requirements of Barrows and Gilts of Different Lean Growth Rates and Allowed Feed Ad Libitum (90 percent dry matter), 113

10-4

Daily Amino Acid Requirements of Barrows and Gilts of Different Lean Growth Rates and Allowed Feed Ad Libitum (90 percent dry matter), 114

10-5

Dietary Mineral, Vitamin, and Fatty Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90 percent dry matter), 115

xv

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

xvi

Tables and Figures 10-6

Daily Mineral, Vitamin, and Fatty Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90 percent dry matter), 116

10-7

Dietary Amino Acid Requirements of Gestating Sows (90 percent dry matter), 117

10-8

Daily Amino Acid Requirements of Gestating Sows (90 percent dry matter), 118

10-9

Dietary Amino Acid Requirements of Lactating Sows (90 percent dry matter), 119

10-10 Daily Amino Acid Requirements of Lactating Sows (90 percent dry matter), 120 10-11 Dietary Mineral, Vitamin, and Fatty Acid Requirements of Gestating and Lactating Sows (90 percent dry matter), 121 10-12 Daily Mineral, Vitamin, and Fatty Acid Requirements of Gestating and Lactating Sows (90 percent dry matter), 122 10-13 Dietary and Daily Amino Acid, Mineral, Vitamin, and Fatty Acid Requirements of Sexually Active Boars (90 percent dry matter), 123 11-1

Chemical Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis), 126

11-2

Mineral Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis), 128

11-3

Vitamin Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis), 130

11-4

Amino Acid Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis), 132

11-5

Apparent Ileal Digestibilities of Amino Acids in Some Feed Ingredients Commonly Used for Swine, 134

11-6

True Ileal Digestibility of Amino Acids in Some Feed Ingredients Commonly Used for Swine, 136

11-7

Coefficients for Estimation of Amino Acids from Crude Protein Content of Feed Ingredients, 138

11-8

Mineral Concentrations in Macro Mineral Sources (data on as-fed basis), 139

11-9

Inorganic Sources and Estimated Bioavailabilities of Trace Minerals, 140

11-10 Characteristics and Energy Values of Various Sources of Fats and Oils (data on asfed basis), 141 11-11 Chemical Composition of Some Purified Feed Ingredients Commonly Used for Swine Research (data on as-fed basis), 141 FIGURES 1-1

Digestible energy intake of growing-finishing pigs as an asymptotic function of body weight. Based on research conducted before 1983 and involving 8,072 observations of 1,390 pens of pigs fed nutritionally adequate corn–soybean meal diets (National Research Council, 1987), 11

2-1

Lysine requirements of starting, growing, and finishing pigs in research published since 1985. Each block represents an estimated requirement (total lysine basis) plotted against the mean body weight of the pigs in the experiment (final body weight minus initial body weight divided by 2). The line represents an estimate of the lysine requirement (total lysine basis), 19

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Tables and Figures

xvii

3-1

Potential whole body protein accretion rate of pigs of high-medium lean growth rate with a carcass fat-free lean gain averaging 325 g/day from 20 to 120 kg body weight (default equation of the model). The lean growth rate of 325 g/day is converted to a mean whole-body protein accretion rate of 127.5 g/day (325/2.55 4 127.5), 32

3-2

Potential whole body protein accretion rates of pigs of medium, high-medium, and high lean growth rates with carcass fat-free lean gains averaging 300, 325, and 350 g/day from 20 to 120 kg body weight (default equation of the model), 33

3-3

Estimated daily digestible energy (DE) intakes of barrows, gilts, and a 1:1 ratio of barrows to gilts consuming feed on an ad libitum basis from 20 to 120 kg body weight (default equation of the model), 33

3-4

Relationship of whole body protein gain and digestible energy intake in pigs from 5 to 150 kg body weight, 34

3-5

Relationship of daily whole body protein deposition and daily intake of true ileal digestible lysine above maintenance. Based on data from 18 experiments and adapted from a summary by Kerr (1993), 34

3-6

Daily lysine requirement (true ileal digestible basis) of pigs with a mean lean growth rate (carcass fat-free basis) of 325 g/day from 20 to 120 kg body weight as estimated by the model using default equations, 35

3-7

Dietary lysine requirement (%, true ileal digestible basis) of pigs with a mean lean growth rate (carcass fat-free basis) of 325 g/day from 20 to 120 kg body weight as estimated by the model using default equations, 36

3-8

Dietary lysine requirements (%) of pigs of medium, high-medium, and high lean growth rates with carcass fat-free lean gains averaging 300, 325, and 350 g/day from 20 to 120 kg body weight as estimated by the model using default equations. The requirements are for total lysine, assuming a corn–soybean meal mixture, 37

3-9

Relation of litter growth rate to dietary apparent ileal digestible lysine intake by lactating sows, 40

3-10

Dietary lysine requirement (%) of pigs from 3 to 20 kg body weight using the default equation of the model (total basis, assuming a corn–soybean meal diet), 41

3-11

Estimated daily feed intake of pigs from 3 to 20 kg and from 20 to 120 kg body weight based on the default equations for digestible energy intake in the model divided by the digestible energy concentration of the diet (3,400 kcal/kg), 41

3-12

Estimated dietary calcium requirement (%) of pigs from 3 to 120 kg body weight using the generalized exponential equation in the model, 42

3-13

Estimated dietary riboflavin requirement (mg/kg) of pigs from 3 to 120 kg body weight using the generalized exponential equation in the model, 42

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirements of Swine Tenth Revised Edition, 1998

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Overview

Energy, amino acids, minerals, vitamins, and water are needed by pigs for body maintenance, growth, reproduction, and lactation. Synthesis of muscle and adipose tissue, bone, hair, skin, and other body components, resulting in accretion of water, protein, lipid, and ash, is dependent upon an adequate dietary supply of nutrients. Pigs must be provided these essential nutrients in adequate amounts and in forms that are palatable and efficiently utilized in order for optimal growth, reproduction, and lactation to occur. Since 1944, the National Research Council has published nine editions of Nutrient Requirements of Swine. This publication has guided nutritionists and other professionals in academia and the swine and feed industries in developing and implementing nutritional and feeding programs for swine. This tenth edition continues that tradition, but the format of this edition is quite different from that of previous ones. The text has been expanded with new sections that address contemporary issues, and the tables are more comprehensive. A new approach using integrated mathematical equations (models) was utilized to generate estimates of energy and amino acid requirements, and a computer program and software allow the user to create tables of nutrient requirements for swine of a specific body weight and level of productivity. The first chapter deals with energy and reviews new information on digestible energy (DE), metabolizable energy (ME), and net energy (NE) requirements of swine. Equations for predicting DE, ME, and NE from chemical components are presented. New information on factors affecting energy requirements of swine also is included in this chapter. In the chapter on proteins and amino acids (Chapter 2), much of the discussion relates to lysine, the first limiting amino acid in most diets for pigs, and to new information on lysine requirements. The concept of ‘‘ideal ratios’’ of essential amino acids to lysine for maximum lean tissue

synthesis in growing pigs and optimal productivity in gestating and lactating sows is described. Discussion of bioavailability of amino acids, on a true and apparent ileal digestibility basis, has been expanded. A section on amino acid requirements of boars is now included. Chapter 3 addresses the use of mathematical models to estimate energy and amino acid requirements of swine. This chapter describes the modeling approach that was taken by the subcommittee to generate the amino acid requirements of growing-finishing pigs from 20 to 120 kg body weight and for gestating and lactating sows. The growth model is based on the pig’s lean growth rate and it estimates the daily true ileal digestible lysine needed to support maximum protein accretion at a given body weight. The dietary lysine requirement is then estimated based on the pig’s daily feed intake, which, in turn, is based on body weight, gender, environmental conditions, and DE concentration of the diet. Estimates of the requirements for other essential amino acids are based on the ideal ratio of each to lysine for maintenance and protein accretion. The gestation model estimates the energy and amino acid requirements of sows based on their breeding weight, targeted gestational weight gain, and litter size. The lactation model estimates requirements based on the sow’s postfarrowing weight, lactational weight change, and daily litter weight gain, a reflection of the amount of milk production. Chapters on minerals and vitamins were updated with results from research studies reported since the previous edition. Chromium is recognized as an essential trace mineral for swine. The sodium and chlorine requirements of the young pig and the manganese requirement of the gestating and lactating sow were increased, and new information on the bioavailability of minerals is presented. The vitamin E and folacin requirements of gestating and lactating sows were increased, based on new research information.

1

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

2

Overview

The chapter on water was expanded. In addition to more discussion of the water requirements of all classes of swine, this section also reviews factors that affect quality of drinking water. The chapter on nonnutritive feed additives was expanded to include antimicrobial agents and other feed additives, including anthelmintics, microbial supplements, oligosaccharides, enzymes, acidifiers, flavors, odor control agents, antioxidants, pellet binders, flow agents, mineral supplements, and carcass modifiers. A new chapter that addresses nutrient excretion was added to the tenth edition. This chapter discusses the potential environmental impact of excessive excretion of nutrients, particularly nitrogen and phosphorus, and addresses means of reducing excretion of these potential environmental pollutants by dietary manipulation. The feed ingredient composition data have been updated and greatly expanded, with 23 additional ingredients added to the tables, for a total of 79 feed ingredients. Net energy, neutral- and acid-detergent fiber (NDF, ADF), and beta-carotene concentrations of feedstuffs were added, and crude fiber was deleted. Vitamin E levels in feedstuffs were modified to include only those assayed by high-performance liquid chromatography. New tables that give estimates of apparent and true digestible coefficients for the amino acids in feedstuffs are now included. Other new tables give the fatty acid composition of fat sources and estimates of the four most limiting amino acids in feedstuffs based on their crude protein content. Finally, the tables of nutrient requirements have been revised and updated. The amino acid requirements are

based on the subcommittee’s assessment of the biological relationships that govern accretion of protein and fat for growth, reproduction, and lactation. The estimates for all nutrients, including amino acids, are based on the best judgment of the subcommittee members following their thorough review of the world’s scientific literature. As in previous editions, the estimated nutrient requirements in this publication are minimum standards without any safety allowances. Therefore, they should not be considered as recommended allowances. Professional nutritionists may choose to increase the levels of some of the more critical nutrients to include ‘‘margins of safety’’ in some circumstances (this comment does not apply to selenium). Another important point is that, for minerals and vitamins, the estimated requirements include the amounts of these nutrients that are present in the natural feedstuffs and are not estimates of amounts of nutrients that should be added to diets. Knowledge of the nutritional needs of swine has expanded considerably since the last revision of this publication. Nevertheless, there is still conflicting, incomplete, or no information for several nutrients at different stages of the life cycle. This is particularly true for many of the vitamins and trace minerals, especially for the very young pig and the gestating and lactating sow. More research is encouraged to expand the knowledge base in these areas. The user of this publication is reminded that knowledge of the principles and assumptions described in the text of this publication is absolutely essential for the proper use of the model and the tables of nutrient requirements.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

1

Energy

CLASSIFICATION OF ENERGY

Energy is produced when organic molecules undergo oxidation. Energy is either released as heat or is trapped in high–energy bonds for subsequent use for the metabolic processes in animals. Energy content in feedstuffs can be expressed as calories (cal), kilocalories (kcal), or megacalories (Mcal) of gross energy (GE), digestible energy (DE), metabolizable energy (ME), or net energy (NE). Energy can also be expressed as joules (J), kilojoules (kJ), or megajoules (MJ) (1 Mcal 4 4.184 MJ; 1 MJ 4 0.239 Mcal; 1 MJ 4 239 kcal). The terms used in this publication to describe energy requirements and energy content of feeds are similar to those defined and extensively discussed in Nutritional Energetics of Domestic Animals and Glossary of Energy Terms (National Research Council, 1981). Whittemore and Morgan (1990), Chwalibog (1991), Ewan (1991), Noblet and Henry (1991), and Hoffmann (1994) have published reviews of energy utilization by swine. Determination of the energy values of feedstuffs for swine is a difficult and tedious task. Originally, energy values were estimated from studies with chicks or were calculated from Total Digestible Nutrients (TDN) (National Research Council, 1971). Since the original direct determinations of energy in feedstuffs for pigs by Diggs et al. (1959, 1965) and Tollett (1961), the database has grown. A summary of energy values of feedstuffs from around the world has been compiled by Ewan (1996). Still, where data are not available by direct means from pig studies, energy concentrations can only be estimated from chemical composition of the feedstuff. Prediction equations that have been used for estimating energy concentrations in feeds are given in the subsequent sections. In all of these equations, the energy and nutrient concentrations are expressed on a dry matter basis.

Gross Energy Gross energy is the energy liberated when a substance is combusted in a bomb calorimeter. The GE concentration of a feed ingredient is dependent on the proportions of carbohydrate, fat, and protein present in the ingredient. Water and minerals contribute no energy; carbohydrates provide 3.7 (glucose) to 4.2 (starch) kcal/g, protein provides 5.6 kcal/g, and fat provides 9.4 kcal/g. If the composition of a feed is known, GE can be predicted fairly accurately. The following relationship was reported by Ewan (1989) for predicting GE (kcal/kg) from ether extract (EE), crude protein (CP), and ash. GE 4 4,143 ` (56 2 % EE) ` (15 2 % CP) 1 (44 2 % Ash), R2 4 0.98 (1-1) Digestible Energy Dietary GE intake minus the GE of the excreted feces is DE. Apparent indigestible energy is a major variable in the evaluation of feed ingredients. Farrell (1978), Agricultural Research Council (1981), and Morgan and Whittemore (1982) suggest that DE is preferable in describing the energy requirements of swine and the energy content of swine feeds, because DE is easily and precisely determined and is, in principle, additive. In addition, DE values are available for most of the commonly used feeds. However, in the conventional scheme of energy utilization, DE is apparent, not true, because fecal metabolic energy is not considered. Chemical composition of feed ingredients is a major determinant of DE, with positive effects of ether extract and negative effects of fiber and ash. The following equations have been reported for predicting DE (kcal/kg) from chemical composition:

3

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

4

Nutrient Requirements of Swine DE 4 1174 ` (0.848 2 GE) ` (2 2 % SCHO) 1 (16 2 % ADF), (1-2) R2 4 0.87; Ewan (1989) DE 4 949 ` (0.789 2 GE) 1 (43 2 % Ash) 1 (41 2 % NDF), R2 4 0.91; Noblet and Perez (1993) (1-3) DE 4 4,151 1 (122 2 % Ash) ` (23 2 % CP) ` (38 2 % EE) 1 (64 2 % CF), (1-4) R2 4 0.89; Noblet and Perez (1993)

in which SCHO is soluble carbohydrate calculated as 100 1 (% CP ` % EE ` % Ash ` % NDF), ADF is acid detergent fiber, NDF is neutral detergent fiber, and CF is crude fiber. Digestibility of dietary energy increases slightly with increased body weight (Noblet and Shi, 1993) because of increased degradation of undigested carbohydrate in the large intestine. Noblet and Shi (1993) proposed that for finishing pigs and particularly sows fed at restricted feed intakes, DE concentrations (kcal/kg) should be corrected by one of the following relationships. DE 4 1,391 ` (0.58 2 DE) ` (23 2 % EE) ` (12.7 2 % CP), R2 4 0.96 (1-5) or, DE 4 1712 ` (1.14 2 DE) ` (33 2 % NDF), R2 4 0.93

(1-6)

Metabolizable Energy The DE minus the GE of gaseous and urinary losses is metabolizable energy (ME). The loss of energy as gas produced in the digestive tract of swine is usually between 0.1 and 3.0 percent of DE (Noblet et al., 1989b; Shi and Noblet, 1993). These amounts are generally ignored because they are small and not easily measured. For most practical swine diets used in North America, ME is 94 to 97 percent of DE, with an average of 96 percent (Farrell, 1979; Agricultural Research Council, 1981). A correction is sometimes made to ME concentrations for nitrogen gained or lost from the body (MEn, Morgan et al., 1975). ME is corrected to nitrogen equilibrium because the energy that is deposited as retained protein cannot be totally recovered by the animal if the amino acids are degraded for energy. This correction to nitrogen equilibrium may be valid for mature animals but is not valid for growing pigs that retain considerable amounts of nitrogen. Therefore, the correction probably is not necessary (Farrell, 1979) or should be made to a constant positive nitrogen retention. The correction factor that is used has been obtained by expressing the GE of urine per gram of urinary nitrogen. For swine, Diggs et al. (1959) used a correction factor of 6.77, Morgan et al. (1975) used 9.17,

and Wu and Ewan (1979) used 7.83 kcal of ME/g of nitrogen to correct for each gram of nitrogen above or below nitrogen equilibrium. This correction is added to the determined ME for pigs in negative nitrogen balance and subtracted when animals are in positive nitrogen balance. If protein is of poor quality or in excess, ME decreases because the amino acids not used for protein synthesis are catabolized and used as a source of energy, and the nitrogen is excreted as urea. Therefore, as the nitrogen content of the urine increases, the energy losses in the urine increase and the ME of the diet decreases. Estimates of ME (kcal/kg) may be calculated from DE (kcal/kg) and CP using one of the following relationships. ME 4 DE 2 (1.012 1 (0.0019 2 % CP)), R2 4 0.91; May and Bell (1971) (1-7) ME 4 DE 2 (0.998 1 (0.002 2 % CP)), R2 4 0.54; Noblet et al. (1989c) (1-8) ME 4 DE 2 (1.003 1 (0.0021 2 % CP)), R2 4 0.48; Noblet and Perez (1993) (1-9) The ME of diets fed to finishing pigs or to sows fed at restricted intakes increases because digestibility is improved. Noblet and Shi (1993) proposed that ME concentrations (kcal/kg) determined with growing pigs (,60 kg) should be adjusted by one of the following relationships for finishing pigs and sows. ME 4 1,107 ` (0.64 2 ME) ` (22.9 2 % EE) (1-10) ` (6.9 2 % CP), R2 4 0.96 or, ME 4 1946 ` (1.17 2 ME) ` (3.15 2 % NDF), R2 4 0.94 (1-11) Net Energy Net energy (NE) is the difference between ME and heat increment (HI). The HI is the amount of heat released because of the energy costs of the digestive and metabolic processes. The energy of the HI is not used for productive processes but can be used to maintain body temperature in cold environments. Net energy, therefore, is the energy that the animal uses for maintenance (NEm) and production (NEp). The energy used for maintenance (NEm) is also dissipated as heat, so that total heat production is the sum of HI and NEm. Evaluation of NE requires the measurement of energy balance or heat production. If energy is required to maintain body temperature or excess activity, NEp is reduced. Although difficult to measure, NE is the best indication of the energy available to an animal for maintenance and production (Noblet et al., 1994). For pigs fed conventional diets and kept at thermoneutral temperatures, the ratio of NE to ME ranged from 0.66

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Energy to 0.75 (Thorbek, 1975; Noblet et al., 1994). Ewan (1976), Phillips and Ewan (1977), and Pals and Ewan (1978) reported the efficiency of ME utilization for energy gain and maintenance (NE) in growing pigs to vary from 27 percent for wheat middlings, to 69 percent for corn, to 75 percent for soybean oil. Noblet et al. (1994) reported efficiencies of energy utilization of 90, 82, 80, 72, and 60 percent for rapeseed oil, cornstarch, sucrose, and mixtures of protein and fiber sources, respectively, for pigs ranging in weight from 45 to 150 kg. Some of the reported relationships between NE (kcal/kg) and chemical composition are as follows: NE 4 328 ` (0.599 2 ME) 1 (15 2 % Ash) 1 (30 2 % ADF), (1-12) R2 4 0.81; Ewan (1989) NE 4 (0.726 2 ME) ` (13.3 2 % EE) ` (3.9 2 % St) 1 (6.7 2 % CP) 1 (8.7 2 % ADF) R2 4 0.97; Noblet et al. (1994) (1-13) NE 4 2,790 ` (41.2 2 % EE) ` (8.1 2 % St) 1 (66.5 2 % Ash) 1 (47.2 2 % ADF), R2 4 0.90; Noblet et al. (1994) (1-14) in which St is starch.

5

(Agricultural Research Council, 1981). Verstegen et al. (1982) estimated that during their growth period, from 25 to 60 kg, pigs needed an additional 25 g of feed/day (80 kcal of ME/day) to compensate for each 1°C below Tc. During the finishing period, from 60 to 100 kg, pigs required an additional 39 g of feed/day (125 kcal of ME/ day) for each 1°C below Tc. For each 1°C below the lower critical temperature (18 to 20°C), there is an increase in heat production of approximately 3.7 to 4.5 kcal of ME/kg of body weight raised to the 0.75 power (BW0.75) (Noblet et al., 1985; Close and Poorman, 1993). The lower critical temperature is reduced by group housing, by use of bedding, and by decreased ventilation rate. For 180-kg sows in normal condition individually housed on concrete, the increase in energy required to maintain body temperature is about 4 percent of maintenance requirement per °C below the lower critical temperature (Verstegen et al., 1987). Between the upper and lower critical temperatures, a zone of thermoneutrality exists where heat production is relatively stable. Environmental temperatures above the critical temperature will reduce feed intake (Ewan, 1976). The National Research Council (1987) suggested that DE intake is reduced by 1.7 percent for each 1°C that the effective ambient temperature of the pig exceeds the upper critical temperature. Here, effective ambient temperature is the temperature the animal experiences.

HEAT PRODUCTION Measurement of total heat production includes the energy associated with HI, the energy required for maintenance, and energy expended in response to changes in the environment. The major environmental factors that influence heat production are temperature and physical activity. Temperature Cold thermogenesis influences energy requirements when the ambient temperature (T, °C) is below the critical temperature (Tc, °C). The critical temperature is the point below which an animal must increase heat production to maintain body temperature. Below T c , the pig must increase its rate of metabolic heat production to maintain homeothermy (National Research Council, 1981). Factors that alter the rate of energy exchange between the animal and its environment will alter Tc (National Research Council, 1981). The energy cost of cold thermogenesis can be described by the following equation: MEHc (kcal ME/day) 4 ((0.313 2 BW) ` 22.71) 2 (Tc 1 T)

Activity Physical activity also influences heat production. Petley and Bayley (1988) measured the heat production of pigs running on a treadmill and reported that heat production of the exercised pigs was 20 percent greater than that of control animals. Close and Poorman (1993) calculated that the additional expenditure of energy by growing pigs for walking was 1.67 kcal of ME/kg of BW for each kilometer. Noblet et al. (1993) measured the increase in heat production associated with standing by sows as 6.5 kcal of ME/ kg of BW 0.75 for ¨each 100 minutes. This figure was similar to reports by Hornicke (1970) of 7.2, by McDonald et al. (1988) of 7.1, by Susenbeth and Menke (1991) of 6.1, and by Cronin et al. (1986) of 7.6 kcal/kg of BW 0.75 for each 100 minutes. Noblet et al. (1993) also determined that the energy cost of consuming feed was 24 to 35 kcal of ME/ kg of feed consumed.

ENERGY REQUIREMENTS (1-15)

where MEHc is energy cost of cold thermogenesis, BW is animal weight in kg, and Tc and T are expressed in °C

Maintenance The ME requirement for maintenance (MEm) includes the needs of all body functions and moderate activity. These

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

6

Nutrient Requirements of Swine

requirements are usually expressed on a metabolic body weight basis, which is defined as body weight raised to the 0.75 power (BW 0.75). Other exponents have been suggested as more appropriate: 0.67 (Heusner, 1982); 0.60 (Noblet et al., 1989b); 0.42 (Noblet et al., 1994). Estimates of the MEm requirement/kg of BW 0.75 vary from 92 to 160 kcal/ day, with most values falling between 100 and 125 kcal/ day. The mean estimate for MEm¨ is 106 kcal of ME/kg of BW 0.75/day (Whittemore, 1976; Bohme et al., 1980; Wenk et al., 1980; Agricultural Research Council, 1981; Noblet and Le Dividich, 1982; Campbell and Dunkin, 1983; Close and Stanier, 1984; McNutt and Ewan, 1984; Gadeken et al., 1985; Noblet et al., 1985), which is equivalent to 110 kcal of DE/kg of BW 0.75. However, Whittemore (1983) suggested that MEm can be more accurately described as: MEm (kcal/day) 4 442 2 Pt 0.78

(1-16)

where Pt is the whole body protein mass in kg. Robles and Ewan (1982) reported daily NE requirements for maintenance (NEm) as 71 kcal/kg of BW 0.75; Just (1982c) reported NEm as 78 kcal/kg of BW 0.75; and Noblet et al. (1994) reported this figure as 86 kcal/kg of BW 0.42. During gestation, 60 to 80 percent of the total energy requirement is used for maintenance. The National Research Council (1988) concluded from the available literature that the daily requirement for maintenance of pregnant sows was 106 kcal of ME or 110 kcal of DE/kg of BW 0.75/day. Noblet et al. (1990), on the basis of recent estimates, concluded that the daily requirement was 105 kcal of ME/kg of BW 0.75 for primiparous and multiparous sows. Beyer et al. (1994) reached a similar conclusion from the literature (103 kcal of ME/kg of BW 0.75/day) for primiparous sows but reported data to indicate an increase from 93 kcal in the first parity to 104 kcal in the second parity and to 113 kcal of ME/kg of BW 0.75 in the fourth parity. Whittemore and Yang (1989) reported the daily requirement as 115 kcal of ME/kg of BW 0.75 from observations over four parities during gestation, lactation, and the interval from weaning to conception. Based on the literature, there seems little justification for altering the value used for growing pigs of 106 kcal of ME/kg of BW 0.75 (or 110 kcal of DE/kg of BW 0.75) for the daily maintenance requirement. Whittemore and Morgan (1990) suggested that the maintenance requirement was proportional to body protein mass (Pt) by the following relationship. MEm (kcal/day) 4 600 2 Pt 0.648

(1-17)

The daily maintenance energy requirement for the lactating sow is presumably also 106 kcal of ME/kg of BW 0.75 (or 110 kcal of DE/kg of BW 0.75) (National Research Council, 1988), which is the same as that for the gestating sow. But some recent reports have suggested that the requirement of the lactating sow may be 5 to 10 percent higher than that of the gestating sow; the higher figure probably

reflects the heat production associated with the synthesis of milk (Noblet and Etienne, 1986, 1987; Burlacu et al., 1986). Noblet et al. (1989a) reported no difference in maintenance requirement among growing boars, barrows, and gilts (112 kcal of ME/kg of BW 0.75). Kemp (1989) reported the maintenance requirement for mature boars as 99 kcal of ME/kg of BW 0.75. McCracken et al. (1991) reported measurement of maintenance requirements of mature boars of 126 kcal of ME/kg of BW 0.75. Although the limited data available may suggest a higher maintenance requirement for boars, the estimate suggested for growing pigs and sows is preferred (106 kcal of ME/kg of BW 0.75 or 110 kcal of DE/kg of BW 0.75). Growth Estimates for the energy costs of protein retention (MEpr) range from 6.8 to 14.0 Mcal of ME/kg, with a mean of 10.6 Mcal of ME/kg (Tess et al., 1984). Literature estimates of the energy costs of fat deposition (MEf) range from 9.5 to 16.3 Mcal of ME/kg, with a mean of 12.5 Mcal of ME/kg (Tess et al., 1984). Although the mean energy costs/kg of protein or fat deposited are approximately equal (Wenk et al., 1980), 1 kg of lean muscle tissue is only 20 to 23 percent protein, whereas 1 kg of adipose tissue is 80 to 95 percent fat. Therefore, the energy cost for muscle tissue production is considerably less than that for fat tissue deposition. Pregnancy The feed and energy requirements of the pregnant sow will vary with body weight, target body weight gain during pregnancy, and other management and environmental parameters. The Agricultural Research Council (1981), Cole (1982), Seerley and Ewan (1983), and Aherne and Kirkwood (1985) reviewed the effects of energy intake during gestation on sow weight gain and reproductive performance. Aherne and Kirkwood (1985) and Williams et al. (1985) suggested that sows should be fed and managed so that they gain 25 kg of maternal tissues throughout pregnancy for at least the first three or four parities. The weight of the placenta and other products of conception should be approximately 20 kg, for a total of 45 kg of gestational weight gain of the sow (Verstegen et al., 1987; Noblet et al., 1990). In general, an increase in the energy intake of the pregnant sow above 6.0 Mcal of ME/day will increase maternal weight gain but will not significantly affect litter size at parturition (Elsley, 1973; Agricultural Research Council, 1981). Whittemore et al. (1984) reported that gestation feed intakes between 1.7 and 2.3 kg/day of sows maintained for five parities had no significant effect on the total number of pigs born. Sows receiving the lowest level of feed did

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Energy have a higher overall culling rate, however. The majority of experiments on this topic have demonstrated that pig birth weights progressively increase when sow feed or energy intake increases during pregnancy. However, a birth weight increase with a maternal feed intake of more than 6.0 Mcal of ME/day is seldom significant (Libal and Wahlstrom, 1977; Henry and Etienne, 1978; Agricultural Research Council, 1981). Increasing feed intake during early gestation does not affect the number of pigs born (den Hartog and van Kempen, 1980; Toplis et al., 1983). High levels of feed intake (. 2.5 kg/day) during the first three days after mating reduced embryo survival in gilts by about 5 percent in one study (Aherne and Williams, 1992) and by 15 percent in another (Dyck et al., 1980), but the reduction in survival does not consistently result in reduced litter size. Elsley et al. (1971) and Cromwell et al. (1980, 1989) demonstrated that the pattern of feed intake during pregnancy was less important in influencing sow performance than the total amount of feed given to the sows. Increasing feed intake in late gestation may increase the average birth weight of pigs (Hillyer and Phillips, 1980; Cromwell et al., 1982). Cromwell et al. (1989) also reported that by increasing feed intake 1.36 kg during the last 23 days of pregnancy, pig weight increased at birth by 40 g and at 21 days of age by 170 g. Weldon et al. (1991) reported that increased energy intake (5.76 to 10.5 Mcal of ME/day) of gilts from day 75 to 105 of pregnancy reduced mammary cell numbers and suggested that milk production may be reduced. Pregnant sows offered feed ad libitum will consume more energy during gestation than required for maintenance and growth of the conceptus tissue, thus resulting in an increase in deposition of body fat and protein. As energy intake and weight gain during pregnancy increase, energy intake during lactation decreases and weight ´loss during lactation increases (Salmon-Legagneur and Rerat, 1962; Baker et al., 1969; Brooks and Smith, 1980; O’Grady, 1980; Cole, 1982; Williams et al., 1985; Weldon et al., 1994). Therefore, it is desirable to limit energy intake during pregnancy to control weight gain. The daily energy requirements for pregnancy include the costs of maintenance, energy required for the deposition of protein and fat in the maternal tissue, and energy requirements of the conceptus. Weight gain during pregnancy is a sum of maternal protein and fat deposition and the gain of the products of conception. Beyer et al. (1994) reported from a comparative slaughter experiment that the total weight gain of the uterus, uterine fluids, products of conception, and mammary tissue was 22.8 kg for sows fed three levels of energy during the first, second, or fourth parity. Assuming a litter size of 10 pigs, this equates to 2.28 kg per pig. The weight gain of protein was 2.46 kg and of fat was 0.46 kg. Total energy gain was 19.94 Mcal. Total maternal weight gain

7

was dependent on the amount of energy consumed. They found that there was an obligatory deposition of about 20 Mcal of NE due to pregnancy, or 174 kcal of NE/day. Assuming an efficiency of utilization of ME for NE of 0.486 (Noblet and Etienne, 1987), the energy requirement for pregnancy would be 358 kcal of ME/day. Additional energy above the maintenance and pregnancy requirement would be used for maternal gain, presumably with the same efficiency as for growth.

Lactation The long-term reproductive efficiency of the sow is best served by minimizing weight loss during lactation (Dourmad et al., 1994). Such a strategy requires only minimal restoration of weight in the next pregnancy. The daily energy requirements during lactation include a requirement for maintenance (MEm) and a requirement for milk production. The energy requirement for milk production can be estimated from the growth rate of the suckling pig and the number of pigs in the litter (Noblet and Etienne, 1989): Milk Energy 4 (4.92 2 ADG 2 pigs) 1 (90 2 pigs)

(1-18)

in which milk energy is in kcal of GE/day, ADG is the growth rate of the suckling pig averaged over the lactation period (g/day), and pigs is the number of pigs in the litter. Assuming that the efficiency of conversion of dietary energy to milk energy is 0.72 (Noblet and Etienne, 1987), the relationship is as described below. ME for Milk 4 (6.83 2 ADG 2 pigs) 1 (125 2 pigs)

(1-19)

If dietary energy intake is not adequate to meet the demands of maintenance and milk production, tissue will be mobilized to provide the necessary nutrients for milk production. Noblet and Etienne (1987) concluded that the efficiency of conversion of tissue energy to milk energy is 0.88; this figure suggests that the major source of energy used is fat.

Developing Boars and Gilts Developing boars and gilts should be given ad libitum access to diets until selected as breeding animals at about 100 kg BW to allow evaluation of the potential growth rate and lean gain. After the animals are selected for the breeding herd, energy intake should be restricted to achieve the desired weight at the time the animals are used for breeding (Wahlstrom, 1991).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

8

Nutrient Requirements of Swine

Sexually Active Boars The energy requirement of the working boar is the sum of the energy required for maintenance, mating activity, semen production, and growth. Kemp (1989) reported that the heat production associated with the collection of semen when mounting a dummy sow was 4.3 kcal of DE/kg of BW 0.75. Close and Roberts (1993) estimated the energy required for semen production from the average energy content of each ejaculation (62 kcal of DE) and an estimate of the efficiency of energy utilization (0.60). The energy required was 103 kcal of DE per ejaculation.

ENERGY SOURCES Sugars and Starch Satisfactory survival and growth rates of pigs fed diets containing high levels of different sugars suggest that glucose and lactose are the sugars most effectively utilized by pigs less than 7 days old (Kidder and Manners, 1978; Sambrook, 1979). Pigs less than 7 days old fed diets containing fructose or sucrose develop severe diarrhea, weight loss, and high mortality (Becker and Terrill, 1954; Aherne et al., 1969). After pigs reach 7 to 10 days of age, they can utilize fructose and sucrose. Starch is the main carbohydrate and energy source in most diets fed to pigs. However, pigs less than 2 to 3 weeks old fed diets containing large amounts of starch do not grow as well as pigs fed diets in which glucose, lactose, or sucrose is the carbohydrate source. The poor growth was attributed to insufficient pancreatic amylase and intestinal disaccharidases (Cunningham, 1959; Sewell and Maxwell, 1966). After pigs are 2 or 3 weeks old, their digestive enzyme systems can digest cereal starch more efficiently. Pigs can then be fed starch- or cereal-based diets (Becker and Terrill, 1954; Cunningham, 1959; Sewell and Maxwell, 1966). Nonstarch Polysaccharides Crude fiber determination is an imprecise analytical procedure. Cellulose, hemicellulose, and lignin in crude fiber are 50 to 80 percent, 20 percent, and 10 to 50 percent, respectively, for typical feedstuffs (Van Soest and McQueen, 1973). In view of the diverse composition of fiber, methods have been developed to quantify fiber based on solubility. Neutral detergent fiber (NDF) is an estimate of the total plant cell wall, which consists primarily of cellulose, hemicellulose, and lignin (Goering and Van Soest, 1970). Acid detergent fiber (ADF) is an estimate of cellulose and lignin. The difference between NDF and ADF is the estimated hemicellulose content of a feed sample (Goering and Van Soest, 1970).

The addition of fiber (crude fiber, NDF, ADF) to swine diets decreases the DE concentration of the diet (King and Taverner, 1975; DeGoey and Ewan, 1975; Kennelly et al., 1978; Kennelly and Aherne, 1980b). Increased feed intake generally results as the pig attempts to maintain DE intake (Baird et al., 1975; Agricultural Research Council, 1981; Low, 1985). When dietary crude fiber exceeds 10 to 15 percent of the diet, however, feed intake may be depressed because of excessive bulk or reduced palatability (Braude, 1967). Low-energy (high-fiber) diets will support growth rates equal to those of pigs fed higher-energy diets during periods of low environmental temperatures, but diets of this type usually depress the growth rate during periods of high temperatures (Coffey et al., 1982; Stahly, 1984). Utilization of fiber by nonruminants has been shown to vary considerably, depending on the fiber source (Bell, 1960; Nehring and Uhlemann, 1972; Laplace and Lebas, 1981), degree of lignification (Forbes and Hamilton, 1952), level of inclusion (Farrell and Johnson, 1970; Just, 1979), and extent of processing (Saunders et al., 1969; McNab, 1975). Fiber utilization is also influenced by the physical and chemical composition of the total diet (Schneider and Lucas, 1950; Myer et al., 1975), level of feeding (Cunningham et al., 1962), age and weight of the animal (Zivkovic and Bowland, 1970), adaptation to the fiber source (Pollman et al., 1979), and individual variation among pigs (Keys et al., 1970; Farrell, 1973; King and Taverner, 1975). When these factors are considered, it is not surprising that the digestibility of ´fiber has been shown to vary between 0 and 97 percent (Rerat, 1978) and that the literature contains conflicting reports about the effects of fiber on the digestibility of nutrients. Just (1982a) reported that an increase in 1 percent of dietary crude fiber depressed digestibility of gross energy by approximately 3.5 percent. Fibrous components of the diet are poorly digested in the small intestine and provide substrates for microbial fermentation in the large intestine. The principal end products of microbial fermentation in the large intestine are volatile fatty acids (VFA). The caloric contribution of VFA to swine has been estimated at values ranging from about 5 to 28 percent of the maintenance energy requirement, depending on the level and frequency of feeding and the fiber level of the diet (Friend et al., 1964; Farrell and Johnson, 1970; Imoto and Namioka, 1978; Kim et al., 1978; Kass et al., 1980; Kennelly et al., 1981). Energy derived from fermentation in the large intestine is utilized with lower efficiency than energy digested in the small intestine (52 versus 76 percent [Noblet et al., 1994]; 57 versus 74 percent [Hoffmann et al., 1990]). There is disagreement concerning the influence of fiber on protein digestibility. Several reports suggest that when the source of fiber does not contribute significant amounts of protein to the diet, then an increase in the level of fiber

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Energy does not affect protein digestibility significantly (Gouwens, 1966; Friend, 1970; Eggum, 1973; Kennelly and Aherne, 1980a). Other researchers have observed, however, that an increase in the dietary level of fiber decreases protein digestibility (Pond et al., 1962; Cole et al., 1967; Kass et al., 1980; Just et al., 1983; Frank et al., 1983; Noblet and Perez, 1993). Lipids The term ‘‘lipid’’ includes both fats and oils. Originally, linoleic and arachidonic acids were both identified as essential fatty acids (EFA) that must be supplied in the diet (Cunnane, 1984). Now it is recognized that these fatty acids are members of N–6 series of EFA and that arachidonic acid can be derived in vivo from linoleic acid. It is difficult to produce overt signs of an EFA deficiency in pigs. Enser (1984) has reported normal growth in pigs from weaning to slaughter weight when they are fed diets containing only 0.1 percent linoleic acid. The Agricultural Research Council (1981) suggested the EFA requirements are 3.0 percent of dietary DE for pigs up to 30 kg and 1.5 percent of dietary DE from 30 to 90 kg. These are equivalent to about 1.2 and 0.6 percent of the diet. Christensen (1985) reported that for maximum performance and efficiency of feed utilization, pigs weaned at 5 weeks of age and raised to 100 kg BW require a dietary lineoleic acid of 0.2 percent of GE, or about 0.1 percent of the diet. This level of linoleic acid is usually present in diets based on commonly used cereal grains and protein supplements. In addition to EFA of the N–6 series, pigs probably require EFA of the N–3 series. However, practical diets also contain adequate amounts of these EFA. Therefore, the main concern is the use of lipids as an energy source. Energy concentrations of selected fats are presented in Chapter 11 (Table 11-10). The value of adding fat to the diets of weanling pigs is uncertain. Pettigrew and Moser (1991) summarized data involving 92 comparisons of fat additions for pigs from 5 to 20 kg. In this weight range, addition of fat reduced growth rate and feed intake while it improved gain-to-feed ratio. The response of growth rate was small (0.01 kg) and variable, with similar numbers of positive (37) and negative (38) responses. Inconsistent responses to added fat may be a result of a number of factors, including the age of the pig at the start of the experiment, the amount of fat added, the type of fat, and the method by which the fat was added. Pettigrew and Moser (1991) reported responses for studies in which a constant protein-to-energy ratio was maintained and found no response in growth rate, a reduction in feed intake, and an improvement in gain-to-feed ratio when fat was added. These data suggest that there is an optimal proteinto-energy ratio for young pigs. Consequently, nutrient

9

requirements often are expressed as the amount per Mcal of DE (Agricultural Research Council, 1981). Such an expression assumes that the optimal nutrient-to-energy ratio for maintenance is the same as for a high level of production. However, this assumption is probably not fully correct because the relative maintenance and gain requirements for specific nutrients probably differ from those for energy. Hence, the ratio will change, usually decreasing as the rate of production or body weight increases. The concept of a fixed optimal protein-to-energy ratio is not supported by the results of several experiments; in these, fat added to diets containing high levels of protein and other nutrients depressed the rate and efficiency of gain (Crampton and Ness, 1954; Smith and Lucas, 1956; Peo et al., 1957; Crampton et al., 1960). Clawson et al. (1962) found little correlation between rate or efficiency of gain and the protein-to-energy ratios. Tribble et al. (1979) and Lewis et al. (1980) reported that the addition of fat to the diet did not influence the lysine requirement of starter pigs fed sorghum- or corn-based diets. Cuaron et al. (1981) reported that protein-to-energy ratios within the range of 53 to 71 g of protein/Mcal of DE did not significantly influence the performance of starter pigs. For growing-finishing swine (20 to 100 kg), the summary by Pettigrew and Moser (1991) indicated consistent improvement in growth rate, reduction in feed intake, improvement in gain-to-feed ratio, but an increase in backfat thickness in response to addition of fat to swine diets. Chiba et al. (1991) reported that a ratio of 3.0 g of lysine (or 49 g of balanced protein) per Mcal of DE was necessary to maximize the beneficial effects of fat addition to diets. The digestibility of the dietary fat, quantity of ME and fat consumed, and environmental temperature in which pigs are housed influence the nutritional value of fat as an energy source for pigs (Stahly, 1984). In general, the substitution of fat for carbohydrate calories in a diet for pigs maintained in a thermoneutral environment increases growth rate and decreases the ME required per unit of body weight gain. But for pigs housed in a warm environment, voluntary ME intake increases by 0.2 to 0.6 percent for each additional 1 percent of fat added to the diet. This increase is because the heat increment of fat is less than that of carbohydrate (Stahly, 1984). The age of the pig, chain length of the fatty acids in the fat, free fatty acid concentration, and unsaturated-tosaturated (U:S) fatty acid ratio influence the apparent digestibility of fat (Stahly, 1984). Dietary fat digestibility is low in the weaned pig and improves as the pig grows. The apparent digestibility of short- or medium-chain fatty acids (14 carbons or less) is high (80 to 95 percent), regardless of the dietary ratio of U:S fatty acids (Stahly, 1984). Powles et al. (1995) summarized a series of studies and reported a curvilinear increase in the digestibility of fat as the ratio of U:S fatty acids increased from 1 to 4. They

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

10

Nutrient Requirements of Swine

also reported a linear decrease in digestibility as free fatty acid concentrations increased from 100 to 800 g/kg of fat. Apparent fat digestibility decreases by 1.3 to 1.5 percent for each additional 1 percent of crude fiber in the diet (Just, 1982a,b,c). Evidence suggests that the addition of fat to the diets of sows during late gestation or lactation increases the milk yield, fat content of colostrum and milk, and survival of pigs from birth to weaning, especially for lightweight pigs (Moser and Lewis, 1980; Coffey et al., 1982; Seerley, 1984; Pettigrew and Moser, 1991). Improvements in survival of pigs from birth to weaning were dependent on the total amount of fat the sow consumed before farrowing (. 1,000 g) and the birth-to-weaning survival of the control groups (, 80 percent). Fat supplementation can also reduce sow weight loss during lactation and decrease the interval from weaning to mating (Moser and Lewis, 1980; Pettigrew, 1981; Cox et al., 1983; Seerley, 1984; Moser et al., 1985; Shurson et al., 1986; Pettigrew and Moser, 1991).

VOLUNTARY FEED INTAKE The control of feed intake is influenced by a number of factors in the following groups: ● Physiological factors, including genetics, neural and hormonal mechanisms, and sensory factors, including olfaction and taste (Baldwin, 1985; Fowler, 1985; National Research Council, 1987); ● Environmental factors, including environmental temperature, humidity, air movement, feeder design and location, number of pigs per pen, and available space per pig (National Research Council, 1987); and ● Dietary factors, including deficiencies or excesses of nutrients, energy density, antibiotics, flavors, feed processing, and availability and quantity of water (Agricultural Research Council, 1981; Fowler, 1985; National Research Council, 1987).

The factors that affect feed intake have been extensively reviewed in Predicting Feed Intake of Major Food-Producing Animals (National Research Council, 1987). These values are for pigs allowed ad libitum access to a balanced corn–soybean meal diet. If the feed intake is restricted, as it sometimes is for gilts and boars used for breeding, the daily nutrient (but not energy) intakes must be maintained at least at the levels suggested for market pigs. To accomplish this, the nutrient-to-energy ratio of the diet must be increased. Voluntary energy intake formulas for various classes of swine are presented below.

Suckling Pigs According to the National Research Council (1987), the DE intake of creep feed by the suckling pig can be expressed by the following relationship: DE intake (kcal/day) 4 1151.7 ` (11.2 2 day), R2 4 0.72

(1-20)

where day is age of the pig. The consumption of dry feed is not predicted until pigs are 13.5 days old.

Weanling Pigs Based on a review of the literature, the National Research Council (1987) concluded that feed intake increases linearly during the postweaning period except for the first 24 hours after weaning, when little or no feed is consumed. Estimates of this rate of increase in feed intake range from 17 to 23 g/day for corn–soybean meal diets containing 3,200 kcal of DE/kg of feed. These data could be described by the following equation: DE intake (kcal/day) 4 11,531 ` (455.5 2 BW) 1 (9.46 2 BW 2), R2 4 0.92

(1-21)

which describes the relationship of BW to the DE intake of the 5- to 15-kg pig.

Growing-Finishing Pigs When growing-finishing pigs weighing 15 to 110 kg are allowed to consume feed ad libitum daily, the energy content of the diet generally controls the amount consumed (Agricultural Research Council, 1981; Cole, 1984; Chiba et al., 1991). Pigs will compensate for decreases or increases in the energy density of the diet by increasing or decreasing their feed intake (Owen and Ridgeman, 1967, 1968; Cole et al., 1968). Within limits, this compensation normalizes energy intake. However, voluntary feed intake varies considerably from day to day and among individual pigs (Frank et al., 1983). For pigs allowed ad libitum access to feed, energy intake is generally about 3 to 4 times the maintenance energy requirement. The National Research Council (1987) described feed intake for pigs that weigh from 15 to 110 kg by an asymptotic relationship to body weight (Figure 1-1). DE intake (kcal/day) 4 13,162 2 (1 1 e10.0176BW)

(1-22)

This equation is similar to a relationship reported by the Agricultural Research Council (1981).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Energy 14,000

Daily DE Intake (kcal)

12,000 10,000 8,000 6,000 -0.0176X

Y = 13,162 x (1-e

4,000

)

2,000 0 0

10

20

30

40

50

60

70

80

90

100 110 120 13

Body Weight (kg)

FIGURE 1-1 Digestible energy intake by growing-finishing pigs as an asymptotic function of body weight. Based on research conducted before 1983 and involving 8,072 observations of 1,390 pens of pigs fed nutritionally adequate corn–soybean meal diets (National Research Council, 1987).

Sows Because feed intake is restricted during gestation, predictions of DE intake are not appropriate. For lactating sows, however, voluntary energy intake responds quadratically, as indicated by the following relationship: DE intake (Mcal/day) 4 13 ` (0.596 2 days) 1 (0.0172 2 days2)

(1-23)

where days is day postfarrowing (National Research Council, 1987). O’Grady et al. (1985) summarized feed intake during lactation from 3,559 sows and observed that feed intake increased with parity, number of pigs nursed, and lactation length but decreased with increased gestation weight gain.

REFERENCES Agricultural Research Council. 1981. The Nutrient Requirements of Pigs: Technical Review. Rev. ed. Slough, England. Commonwealth Agricultural Bureaux. xxii, 307 pp. Aherne, F. X., and R. N. Kirkwood. 1985. Nutrition and sow prolificacy. J. Reprod. Fertil. Suppl. 33:169–183. Aherne, F. X., and I. H. Williams. 1992. Nutrition for optimizing breeding herd performance. Vet. Clinics of N. America: Food Anim. Practice 8:589–608. Aherne, F. X., V. W. Hays, R. C. Ewan, and V. C. Speer. 1969. Absorption and utilization of sugars by the baby pig. J. Anim. Sci. 29:444–450. Baker, D. H., D. E. Becker, H. W. Norton, C. E. Sasse, A. H. Jensen, and B. G. Harmon. 1969. Reproductive performance and progeny development in swine as influenced by feed intake during pregnancy. J. Nutr. 97:489–495. Baird, D. M., H. C. McCampbell, and J. R. Allison. 1975. Effect of level of crude fiber, protein and bulk in diets for finishing hogs. J. Anim. Sci. 41:1039–1047. Baldwin, B. A. 1985. Neural and hormonal mechanisms regulating food intake. Proc. Nutr. Soc. 44:303–311.

11

Becker, D. E., and D. E. Terrill. 1954. Various carbohydrates in a semipurified diet for the growing pig. Arch. Biochem. Biophys. 50:399–403. Bell, J. M. 1960. A comparison of fibrous feedstuffs in nonruminant rations. Effects on growth responses, digestibility, rates of passage and ingesta volume. Can. J. Anim. Sci. 40:71–82. Beyer, M. W. Jentsch, L. Hoffmann, R. Schiemann, and M. Klein. 1994. Untersuchungen zum energie- und stickstoffumsatz von graviden und laktierend saun sowie von saugferkeln 4. Mitteilung—Chemische Zusammensetzung und energiegehalt der Konzeptionsprodukte, der reproduktiven Organe und der Lebendmassezunahmmen order abnahmen bei graviden und laktierenden Sauen. Arch. Anim. Nutr. ¨46:7–35. Bohme, H., D. Gadeken, and H. J. Oslage. 1980. Studies on energy costs of protein and fat deposition in early weaned piglets. Landw. Forsch. 33:261–271. Braude, R. 1967. The effect of changes in feeding patterns on the performance of pigs. Proc. Nutr. Soc. 26:163–181. Brooks, P. H., and D. A. Smith. 1980. The effect of mating age on the reproductive performance, food utilization and liveweight change of the female pig. Livestock Prod. Sci. 7:67–78. Burlacu, G., M. Iliescu, and P. Caramida. 1986. Efficiency of feed utilization by pregnant and lactating sows. 2. The influence of isocaloric diets with different protein levels on pregnancy and lactation. Arch. Tierernahr. 36:803–825. Campbell, R. G., and A. C. Dunkin. 1983. The effects of energy intake and dietary protein on nitrogen retention, growth performance, body composition and some aspects of energy metabolism of baby pigs. Br. J. Nutr. 49:221–230. Chiba, L. I., A. J. Lewis, and E. R. Peo, Jr. 1991. Amino acid and energy interrelationships in pigs weighing 20 to 50 kilograms: I. Rate and efficiency of weight gain. J. Anim. Sci. 69:694–707. Christensen, K. 1985. Determination of linoleic acid requirements in slaughter pigs. Res. Rep. No. 577. Copenhagen, Denmark: Beret. Statens Husdyrbrugsforsog. Chwalibog, A. 1991. Energetics of animal production. Acta Agric. Scand. 41:147–160. Clawson, A. J., T. N. Blumer, W. W. G. Smart, Jr., and E. R. Barrick. 1962. Influence of energy-protein ratio on performance and carcass characteristics of swine. J. Anim. Sci. 21:62–68. Close W. H., and P. K. Poorman. 1993. Outdoor pigs—their nutrient requirements, appetite and environmental responses. Pp. 175–196 in Recent Advances in Animal Nutrition, P. C. Garnsworthy, and D. J. A. Cole, eds. Loughborough, U.K.: Nottingham University Press. Close, W. H. and F. G. Roberts. 1993. Nutrition of the working boar. Pp. 21–44 in Recent Advances in Animal Nutrition, W. Haresign, and D. J. A. Cole, eds. Loughborough, U.K.: Nottingham University Press. Close, W. H., and M. W. Stanier. 1984. Effects of plane of nutrition and environmental temperature on the growth and development of the early weaned pig. 2. Energy production. Anim. Prod. 38:221–231. Coffey, M. T., R. W. Seerley, D. W. Funderburke, and H. C. McCampbell. 1982. Effect of heat increment and level of dietary energy and environmental temperature on the performance of growing-finishing swine. J. Anim. Sci. 54:95–105. Cole, D. J. A. 1982. Nutrition and reproduction. Pp. 603–619 in Control of Pig Reproduction, D. J. A. Cole and G. J. Foxcroft, eds. London: Butterworth. Cole, D. J. A. 1984. The nutrient density of pig diets—allowances and appetite. Pp. 301–312 in Fats in Animal Nutrition, J. Wiseman, ed. London: Butterworth. Cole, D. J. A., J. E. Duckworth, and W. Holmes. 1967. Factors affecting voluntary feed intake in pigs. 1. The effect of digestible energy content of the diet on the intake of castrated male pigs housed in holding pens and in metabolism crates. Anim. Prod. 9:141–148.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

12

Nutrient Requirements of Swine

Cole, D. J. A., E. Duckworth, W. Holmes, and A. Cuthbertson. 1968. Factors affecting voluntary feed intake in pigs. 3. The effect of a period of feed restriction, nutrient density of the diet and sex on intake, performance and carcass characteristics. Anim. Prod. 10:345–357. Cox, N. M., J. H. Britt, W. D. Armstrong, and H. D. Alhusen. 1983. Effects of feeding fat and altering weaning schedule on rebreeding in primiparous sows. J. Anim. Sci. 56:21–29. Crampton, E. W., and O. M. Ness. 1954. A meal mixture suitable as the entire ration to be self-fed dry to pigs weaned at ten days of age. J. Anim. Sci. 13:357–364. Crampton, E. W., R. K. Shaw, V. G. MacKay, and D. C. Schad. 1960. Comparative feeding value of common edible fats as measured by the growth of prematurely weaned pups, guinea pigs and swine. J. Nutr. 70:81–90. Cromwell, G. L., T. S. Stahly, and J. R. Overfield. 1980. Evaluation of a constant vs. stepwise feeding pattern for pregnant sows. J. Anim. Sci. 51(Suppl. 1):194 (Abstr.). Cromwell, G. L., T. J. Prince, G. E. Combs, C. V. Maxwell, D. A. Knabe, and D. E. Orr. 1982. Effects of additional feed during late gestation on reproductive performance of sows: A cooperative study. J. Anim. Sci. 55(Suppl. 1):268 (Abstr.). Cromwell, G. L., D. D. Hall, A. J. Clawson, G. E. Combs, D. A. Knabe, C. V. Maxwell, P. R. Noland, D. E. Orr, Jr., and T. J. Prince. 1989. Effects of additional feed during late gestation on reproductive performance of sows: A cooperative study. J. Anim. Sci. 67:3–14. Cronin, G. M., F. M. F. M. Van Tartwijk, W. Van Der Hel, and M. W. A. Verstegen. 1986. The influence of degree of adaptation to tetherhousing by sows in relation to behavior and energy metabolism. Anim. Prod. 42:257–268. Cuaron, J. A., F. X. Aherne, R. A. Easter, A. H. Jensen, and T. F. Park. 1981. Dietary protein–calorie ratios for young pigs. Proc. Illinois Pork Industry Conference. Urbana: Illinois Agricultural Experiment Station. Cunnane, S. C. 1984. Essential fatty-acid/mineral interactions with reference to the pig. Pp. 167–183 in Fats in Animal Nutrition, J. Wiseman, ed. London: Butterworth. Cunningham, H. M. 1959. Digestion of starch and some of its degradation products by newborn pigs. J. Anim. Sci. 18:964–975. Cunningham, H. M., D. W. Friend, and J. W. G. Nicholson. 1962. The effect of age, body weight, feed intake and adaptability of pigs on the digestibility and nutritive value of cellulose. Can. J. Anim. Sci. 42:167–175. DeGoey, L. W., and R. C. Ewan. 1975. Effect of level of intake and diet dilution on energy metabolism in the young pig. J. Anim. Sci. 40:1045–1051. den Hartog, L. A., and G. J. M. van Kempen. 1980. Relation between nutrition and fertility in pigs. Neth. J. Agric. Sci. 28:211–227. Diggs, B. G., D. E. Becker, S. W. Terrill, and A. H. Jensen. 1959. The energy value of various feedstuffs for the young pig. J. Anim. Sci. 18:1492 (Abstr.). Diggs, B. G., D. E. Becker, A. H. Jensen, and H. W. Norton. 1965. Energy value of various feeds for the young pig. J. Anim. Sci. 24:555–558. Dourmad, J.Y., M. Etienne, A. Prunier, and J. Noblet. 1994. The effect of energy and protein intake of sows on their longevity: A review. Livest. Prod. Sci. 40:87–97. Dyck, G. W., W. M. Palmer, and S. Simaraks. 1980. Progesterone and lutenizing hormone concentration in serum of pregnant gilts on different levels of feed consumption. Can. J. Anim. Sci. 60:877–884. Eggum, B. O. 1973. A Study of Certain Factors Influencing Protein Utilization in Rats and Pigs. Ph.D. dissertation. National Institute of Animal Science, Copenhagen, Denmark. Elsley, F. W. H. 1973. Nutrition of the female pig during pregnancy and lactation. Paper presented to Pig Commission. Vienna: European Association of Animal Production.

Elsley, F. W. H., E. V. J. Bathurst, A. G. Bracewell, J. M. M. Cunningham, J. B. Dent, T. L. Dodsworth, R. M. MacPherson, and N. Walker. 1971. The effect of pattern of food intake in pregnancy upon sow productivity. Anim. Prod. 13:257–270. Enser, M. 1984. The chemistry, biochemistry and nutritional importance of animal fats. Pp. 23–51 in Fats in Animal Nutrition, J. Wiseman, ed. London: Butterworth. Ewan, R. C. 1976. Utilization of energy of feed ingredients by young pigs. Proc. Distill. Feed Res. Council. Conf. 31:16–21. Ewan, R. C. 1989. Predicting the energy utilization of diets and feed ingredients by pigs. Pp. 271–274 in Energy Metabolism, European Association of Animal Production Bulletin No. 43, Y. van der Honing and W. H. Close, eds. Pudoc Wageningen, Netherlands. Ewan, R. C. 1991. Energy Utilization in swine nutrition. Pp. 121–132 in Swine Nutrition, E. R. Miller, D. W. Ullrey, and A. J. Lewis, eds. Stoneham, U.K.: Butterworth-Heinemann. Ewan, R. C. 1996. Energy Values of Feed Ingredients. 5th Revised Edition. Ames: Iowa State University. 101 pp. Farrell, D. J. 1973. Digestibility by pigs of the major chemical components of diets high in plant cell-wall constituents. Anim. Prod. 16:43–47. Farrell, D. J. 1978. Metabolizable energy in feeding systems for pigs and poultry. Proc. Aust. Soc. Anim. Prod. 12:62–67. Farrell, D. J. 1979. Energy systems for pigs and poultry: A review. J. Aust. Inst. Agric. Sci. 34:21–34. Farrell, D. J., and K. A. Johnson. 1970. Utilization of cellulose by pigs and its effects on caecal function. Anim. Prod. 14:209–217. Forbes, R. M., and T. S. Hamilton. 1952. The utilization of certain cellulosis materials by swine. J. Anim. Sci. 11:480–490. Fowler, V. R. 1985. The importance of voluntary feed intake in pigs. Proc. Nutr. Soc. 44:347–353. Frank, G. R., F. X. Aherne, and A. H. Jensen. 1983. A study of the relationship between performance and dietary component digestibilities by swine fed different levels of dietary fiber. J. Anim. Sci. 57:645–654. Friend, D. W. 1970. Comparison of some milling products of barley and rye when fed in diets to rats. Can. J. Anim. Sci. 50:345–348. Friend, D. W., J. W. G. Nicholson, and H. M. Cunningham. 1964. Volatile fatty acid and lactic acid content of pig blood. Can. J. Anim. Sci. 44:303–308. ¨ Gadeken, D., H. J. Oslage, and H. Bohme. 1985. Energy requirement for maintenance and energy costs of protein and fat deposition in piglets. Arch. Tierernahr. 35:481–494. Goering, H. K., and P. J. Van Soest. 1970. Forage fiber analysis: Apparatus, reagents, procedures and some applications. Agricultural Handbook 379. Washington, D.C.: U.S. Department of Agriculture. iv, 20 pp. Gouwens, D. W. 1966. Influence of Dietary Protein and Fiber on Fecal Amino Acid Excretion of Swine. M.S. thesis. University of Illinois, Urbana. Henry, Y., and M. Etienne. 1978. Alimentation energetique du porc. J. Rech. Porc. en France 10:119–166. Heusner, A. A. 1982. Energy metabolism and body size. 1. Is the mass exponent of Kleiber’s equation a statistical artifact? Respir. Physiol. 48:1–12. Hillyer, G. M., and P. Phillips. 1980. The effect of increasing feed level to sows and gilts in late pregnancy on subsequent litter size, litter weight and maternal body weight change. Anim. Prod. 30:469 (Abstr.). Hoffmann, L. 1994. Vorschlag fur die einfuhrung eines fur alle nutztierarten nach einheitlichen parametern und prinzipien aufgebauten systems der enmergetischen ¨ futterbewertung. 2. Mitteilung—Energetische Futterbewertung fur Schweine. Arch Anim. Nutr. 46:237–259. Hoffmann, L., W. Jentsch, and R. Schiemann. 1990. Energieumsatzmes¨ sungen bei Vertutterung von Rationen mit Kartof¨ am adulten Schwein ¨ felstarke, Kartoffeln, Ruben, Presschitzeln and Grobfuttermitteln als

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Energy Zulagen zu einer Grundration. 1. Energienumsatz and Energieverwer¨tung. Arch. Anim. Nutr. 40:191–207. Hornicke, H. 1970. Circadian activity rhythms and the energy cost of standing in growing pigs. Pp. 165–168 in Energy Metabolism of Farm Animals, A. Suchurch and C. Wenk, Eds., EAAP No. 13. Juris Druck Verlag, Zurich. Imoto, S., and S. Namioka. 1978. VFA production in the pig’s large intestine. J. Anim. Sci. 47:467–478. Just, A. 1979. Influence of diet composition on site of absorption and efficiency of utilization of metabolizable energy in growing pigs. Pp. 27–30 in Energy Metabolism, European Association of Animal Production Bulletin No. 26, L.E. Mount, ed. Cambridge: Butterworth. Just, A. 1982a. The influence of crude fiber from cereals on the net energy value of diets for growth in pigs. Livest. Prod. Sci. 9:569–580. Just, A. 1982b. The influence of ground barley straw on the net energy value of diets for growth in pigs. Livest. Prod. Sci. 9:717–729. Just, A. 1982c. The net energy value of balanced diets for growing pigs. Livest. Prod. Sci. 8:541–555. Just, A., J. A. Fernandez, and H. Jorgensen. 1983. The net energy value of diets for growth in pigs in relation to the fermentative processes in the digestive tract and the site of absorption of the nutrients. Livest. Prod. Sci. 10:171–186. Kass, M. L., P. J. Van Soest, and W. G. Pond. 1980. Utilization of dietary fiber from alfalfa by growing swine. II. Volatile fatty acid concentrations in and disappearance from the gastrointestinal tract. J. Anim. Sci. 50:192–197. Kemp, B. 1989. Investigations on breeding boars to contribute to a functional feeding strategy. Ph.D. Dissertation, University of Wageningen, The Netherlands. Kennelly, J. J., and F. X. Aherne. 1980a. The effect of fiber addition to diets formulated to contain different levels of energy and protein on growth and carcass quality of swine. Can. J. Anim. Sci. 60:385–393. Kennelly, J. J., and F. X. Aherne. 1980b. The effect of fiber formulated to contain different levels of energy and protein on digestibility coefficients in swine. Can. J. Anim. Sci. 60:717–726. Kennelly, J. J., F. X. Aherne, and A. J. Lewis. 1978. The effects of levels of isolation, or varietal differences in high fiber hull fraction of low glucosinolate rapeseed meals on rat or pig performance. Can. J. Anim. Sci. 58:743–752. Kennelly, J. J., F. X. Aherne, and W. C. Sauer. 1981. Volatile fatty acid production in the hindgut of swine. Can. J. Anim. Sci. 61:349–361. Keys, J. E., Jr., P. J. Van Soest, and E. P. Young. 1970. Effect of increasing cell wall content on the digestibility of hemicellulose and cellulose in swine and rats. J. Anim. Sci. 31:1172–1177. Kidder, D. E., and M. J. Manners. 1978. Digestion of carbohydrates. Pp. 96–149 in Digestion in the Pig. Bath, England: Kington Press. Kim, K. I., D. E. Jewell, N. J. Benevenga, and R. H. Grummer. 1978. The fraction of dietary lactose available for fermentation in the cecum and colon of pigs. J. Anim. Sci. 46:1658–1665. King, R. H., and M. R. Taverner. 1975. Prediction of the digestible energy in pig diets from analyses of fibre contents. Anim. Prod. 21:275–284. Laplace, J. P., and F. Lebas. 1981. Nutritional value of plantex (fiber) in animal feeding. World Rev. Nutr. Diet 37:177–228. Lewis, A. J., E. R. Peo, Jr., B. D. Moser, and T. D. Crenshaw. 1980. Lysine requirements of pigs weighing 5 to 15 kg fed practical diets with and without added fat. J. Anim. Sci. 51:361–366. Libal, G. W., and R. C. Wahlstrom. 1977. Effect of level of feeding during lactation on sow and pig performance. J. Anim. Sci. 41:1524–1525. Low, A. G. 1985. The role of dietary fibre in digestion absorption and metabolism. Proc. 3rd Int. Seminar on Digestive Physiology in the Pig. Report No. 580. Copenhagen, Denmark: Beret. Statens. Husdyrbugsfors. May, R. W., and J. M. Bell. 1971. Digestible and metabolizable energy values of some feeds for the growing pig. Can. J. Anim. Sci. 51:271–278.

13

McCracken, K. J., D. S. Rao, and R. Urquhart. 1991. Feed intake, body composition and energy metabolism of high genetic potential boars from 30 to 340 kg. Pp. 111–114 in Energy Metabolism of Farm Animals, C. Wenk and M. Boessinger, eds. ETH, Zurich. McDonald, T. P., D. D. Jones, J. R. Barret, J. L. Albright, G. E. Miles, J. A. Nienaaber, and G. L. Hahn. 1988. Measuring the heat increment of activity of growing-finishing swine. Trans. Am. Soc. Agric. Eng. 31:1180–1186. McNab, J. M. 1975. Factors affecting the digestibility of nutrients. Proc. Nutr. Soc. 34:5–11. McNutt, S. D., and R. C. Ewan. 1984. Energy utilization of weanling pigs raised under pen conditions. J. Anim. Sci. 59:738–745. Morgan, C. A., and C. T. Whittemore. 1982. Energy evaluation of feeds and compounded diets for pigs. A review. Anim. Feed Sci. Technol. 7:387–400. Morgan, D. J., D. J. A. Cole, and D. Lewis. 1975. Energy values in pig nutrition. I. The relationship between digestible energy, metabolizable energy and total digestible nutrient values of a range of feedstuffs. J. Agric. Sci. (Camb.) 84:7–17. Moser, B. D., and A. J. Lewis. 1980. Adding fat to sow diets. Feedstuffs 52:36–37. Moser, R. L., J. E. Pettigrew, S. G. Cornelius, and H. E. Hanke. 1985. Feed and energy consumption by lactating sows as affected by supplemental dietary fat. Minn. Swine Res. Rep. St. Paul: University of Minnesota Press. Myer, R. O., P. R. Cheeke, and W. H. Kennick. 1975. Utilization of alfalfa protein concentrate by swine. J. Anim. Sci. 40:885–891. National Research Council. 1971. Atlas of Nutritional Data on United States and Canadian Feeds. Washington, D.C.: National Academy of Sciences–National Research Council. National Research Council. 1981. Nutritional Energetics of Domestic Animals and Glossary of Energy Terms. Second revised ed. Washington, D.C.: National Academy Press. 54 pp. National Research Council. 1987. Predicting Feed Intake of FoodProducing Animals. Washington, D.C.: National Academy Press. 85 pp. National Research Council. 1988. Nutrient Requirements of Swine. Ninth Edition. Washington, D.C.: National Academy Press. 93 pp. Nehring, K., and H. Uhlemann. 1972. Feeding value of different byproducts of cereal grain milling. 4. Relationship between the content of structural substances and digestibility. Arch. Tierernahr. 22:59–84. Noblet, J., and J. Le Dividich. 1982. Effect of environmental temperature and feeding level on energy balance traits of early-weaned piglets. Livest. Prod. Sci. 9:619–632. Noblet, J., and M. Etienne. 1986. Effect of energy level in lactating sows on yield and composition of milk and nutrient balance of piglets. J. Anim. Sci. 63:1888–1896. Noblet, J., and M. Etienne. 1987a. Metabolic utilization of energy and maintenance requirements in lactating sows. J. Anim. Sci. 64:774–781. Noblet, J., and M. Etienne. 1987b. Metabolic utilization of energy and maintenance requirements in pregnant sows. Livest. Prod. Sci. 16:243–257. Noblet, J., and M. Etienne. 1989. Estimation of sow milk nutrient output. J. Anim. Sci. 67:3352–3359. Noblet, J., and Y. Henry. 1991. Energy evaluation systems for pig diets. Pp. 87–110 in Manipulating Pig Production III, E. S. Batterham, ed. Australasian Pig Science Association, Attwood, Australia. Noblet, J., and J. M. Perez. 1993. Prediction of digestibility of nutrients and energy values of pig diets from chemical analysis. J. Anim. Sci. 71:3389–3398. Noblet, J., and X. S. Shi. 1993. Comparative digestibility of energy and nutrients in growing pigs fed ad libitum and adult sows at maintenance. Livest. Prod. Sci. 34:137–152. Noblet, J., J. Le Dividich, and T. Bikawa. 1985. Interaction between energy level in the diet and environmental temperature on the utilization of energy in growing pigs. J. Anim. Sci. 61:452–459.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

14

Nutrient Requirements of Swine

Noblet, J., C. Karege, and S. Dubois. 1989a. Influence of sex and genotype on energy utilization in growing pigs. Pp. 57–60 in Energy Metabolism of Farm Animals, Y. van der Honing and W. H. Close, eds. Pudoc Wageningen. Noblet, J., J. Y. Dourmad, J. Le Dividich, and S. Dubois. 1989b. Effect of ambient temperature and addition of straw or alfalfa in the diet on energy metabolism of pregnant sows. Livest. Prod. Sci. 21:309–324. Noblet, J, H. Fortune, S. Dubois, and Y. Henry. 1989c. Nouvelles Bases D’Estimations Des Teneurs en Energie Digestible, Metabolisable et Nette Des Aliments Pour Le Porc. Institut National de la Recherche Agronomique, Paris, France. Noblet, J., J. Y. Dourmad, and M. Etienne. 1990. Energy utilization in pregnant and lactating sows: Modeling of energy requirements. J. Anim. Sci. 68:562–572. Noblet, J., X. S. Shi, and S. Dubois. 1993. Energy cost of standing activity in sows. Livest. Prod. Sci. 34:127–136. Noblet, J., X. S. Shi, and S. Dubois. 1994. Effect of body weight on net energy value of feeds for growing pigs. J. Anim. Sci. 72:645–657. Noblet, J., H. Fortune, X. S. Shi, and S. Dubois. 1994. Prediction of net energy value of feeds for growing pigs. J. Anim. Sci. 72:344–354. O’Grady, J. F. 1980. Energy and protein nutrition of the sow. Pp. 121–131 in Recent Advances in Animal Nutrition, W. Haresign, ed. London: Butterworth. O’Grady, J. F., P. B. Lynch, and P. A. Kearney. 1985. Voluntary feed intake by lactating sows. Livest. Prod. Sci. 12:355–366. Owen, J. B., and W. J. Ridgeman. 1967. The effect of dietary energy content on the voluntary intake of pigs. Anim. Prod. 9:107–113. Owen, J. B., and W. J. Ridgeman. 1968. Further studies on the effect of dietary energy content on the voluntary intake of pigs. Anim. Prod. 10:85–91. Pals, D. A., and R. C. Ewan. 1978. Utilization of the energy of dried whey and wheat middlings by young swine. J. Anim. Sci. 46:402–408. Peo, E. R., Jr., G. C. Ashton, V. C. Speer, and D. V. Catron. 1957. Protein and fat requirements of baby pigs. J. Anim. Sci. 16:885–891. Petley, M. P., and H. S. Bayley. 1988. Exercise and post-exercise energy expenditure in growing pigs. Can. J. Physiol. Pharam. 66:721–730. Pettigrew, J. E., Jr. 1981. Supplemental dietary fat for peripartal sows: A review. J. Anim. Sci. 53:107–117. Pettigrew, J. E., Jr., and R. L. Moser. 1991. Fat in swine nutrition. Pp. 133–146 in Swine Nutrition, E. R. Miller, D. E. Ullrey, and A. J. Lewis, eds. Stoneham, U.K.: Butterworth-Heinemann. Phillips, B. C., and R. C. Ewan. 1977. Utilization of energy of milo and soybean oil by young swine. J. Anim. Sci. 44:990–997. Pollman, D. S., D. M. Danielson, and E. R. Peo, Jr. 1979. Value of highfiber diets for gravid swine. J. Anim. Sci. 48:1385–1393. Pond, W. G., R. S. Lowrey, and J. H. Maner. 1962. Effect of crude fiber level on ration digestibility and performance in growing-finishing swine. J. Anim. Sci. 21:692–696. Powles, J., J. Wiseman, D.J.A. Cole, and S. Jagger. 1995. Prediction of the apparent digestible energy value of fats given to pigs. Anim. ´Sci. 61:149–154. Rerat, A. 1978. Digestion and absorption of carbohydrates and nitrogenous matters in the hind gut of the omnivorous nonruminant animal. J. Anim. Sci. 46:1808–1837. Robles, A., and R. C. Ewan. 1982. Utilization of energy of rice and rice bran by young pigs. J. Anim. Sci. ´ 55:572–577. Salmon-Legagneur, E., and A. Rerat. 1962. Nutrition of the sow during pregnancy. Pp. 207–237 in Nutrition of Pigs and Poultry, J. T. Morgan, and D. Lewis, eds. London: Butterworth. Sambrook, I. E. 1979. Digestion and absorption of carbohydrate and lipid in the stomach and small intestine of the pig. In Technical Bulletin No. 3, Current Concepts of Digestion and Absorption in Pigs, A. G. Low, and I. G. Partridge, eds. Reading, U.K.: National Institute for Research in Dairying.

Saunders, R. M., H. G. Walker, Jr., and G. O. Kohler. 1969. Aleurone cells and the digestibility of wheat mill feeds. Poult. Sci. 48:1497–1503. Schneider, B. H., and H. L. Lucas. 1950. The magnitude of certain sources of variability in digestibility data. J. Anim. Sci. 9:504–512. Seerley, R. W. 1984. The use of fat in sow diets. Pp. 333–352 in Fats in Animal Nutrition, J. Wiseman, ed. London: Butterworth. Seerley, R. W., and R. C. Ewan. 1983. An overview of energy utilization in swine nutrition. J. Anim. Sci. 57(Suppl. 2):300–314. Sewell, R. F., and C. V. Maxwell. 1966. Effects of various sources of carbohydrates in the diet of early weaned pigs. J. Anim. Sci. 25:796–799. Shi, X. S., and J. Noblet. 1993. Digestible and metabolizable energy values of ten feed ingredients in growing pigs fed ad libitum and sows fed at maintenance level: Comparative contribution of the hindgut. Anim. Feed Sci. Tech. 42:223–236. Shurson, G. C., M. G. Hogberg, N. DeFever, S. V. Radecki, and E. R. Miller. 1986. Effects of adding fat to the sow lactation diet on lactation and breeding performance. J. Anim. Sci. 62:672–680. Smith, H., and I. A. M. Lucas. 1956. The early weaning of pigs. 1. The effect upon growth of variations in the protein, fat, sucrose, antibiotic, vitamin and mineral content of diets for pigs of 8 to 25 lb. liveweight and a comparison of wet and dry feeding. J. Agric. Sci. (Camb.) 48:220. Susenbeth, A., and K. H. Menke. 1991. Energy requirement for physical activity in pigs. Pp. 416–419 in Energy Metabolism of Farm Animals, C. Wenk and M. Boessinger, eds. ETH, Zurich. Stahly, T. S. 1984. Use of fats in diets for growing pigs. Pp. 313–331 in Fats in Animal Nutrition, J. Wiseman, ed. London: Butterworth. Tess, M. H., G. E. Dickerson, J. A. Nienaber, J. T. Yen, and C. L. Farrell. 1984. Energy costs of protein and fat deposition in pigs fed ad libitum. J. Anim. Sci. 58:111–122. Thorbek, G. 1975. Studies on energy metabolism in growing pigs. II. Protein and fat gain in growing pigs fed different feed compounds. Efficiency of utilization of metabolizable energy for growth. Research Report No. 424. Copenhagen, Denmark: Beret. Statens. Husdyrbrugsforsog. Tollett, J. T. 1961. The available energy content of feedstuffs for swine. Ph.D. Dissertation. Urbana: University of Illinois. 104 pp. Toplis, P., M. F. J. Ginesi, and A. E. Wrathhall. 1983. The influence of high food levels in early pregnancy on embryo survival in multiparous sows. Anim. Prod. 37:45–48. Tribble, L. F., S. H. Ingram, C. T. Gaskins, and C. B. Ramsey. 1979. Evaluation of added fat and lysine to sorghum-soybean meal diets for swine. J. Anim. Sci. 48:541–546. Van Soest, P. J., and R. W. McQueen. 1973. The chemistry and estimation of fibre. Proc. Nutr. Soc. 32:123–130. Verstegen, M. W. A., H. A. Brandsma, and G. Mateman. 1982. Feed requirement of growing pigs at low environmental temperatures. J. Anim. Sci. 55:88–94. Verstegen, M. W. A., J. M. F. Verhagen, and L. A. den Hartog. 1987. Energy requirements of pigs during pregnancy: A review. Livest. Prod. Sci. 16:75–89. Wahlstrom, R. C. 1991. Feeding developing gilts and boars. Pp. 517–526 in Swine Nutrition, E. R. Miller, D. E. Ullrey, and A. J. Lewis, eds. Stoneham, U.K.: Butterworth-Heinemann. Weldon, W. C., A. J. Thulin, O. A. MacDougald, L. J. Johnston, E. R. Miller, and H. A. Tucker. 1991. Effects of increased dietary energy and protein during late gestation on mammary development in gilts. J. Anim. Sci. 69:194–200. Weldon, W. C., A. J. Lewis, G. F. Louis, J. L. Kovar, M. A. Giesemann, and P. S. Miller. 1994. Postpartum hypophagia in primiparous sows: I. Effects of gestation feeding level on feed intake, feeding behavior, and plasma metabolite concentrations during lactation. J. Anim. Sci. 72:387–394. Wenk, C., H. P. Pfirter, and H. Bickel. 1980. Energetic aspects of feed conversion in growing pigs. Livest. Prod. Sci. 7:483–495.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Energy Whittemore, C. T. 1976. A study of growth responses to nutrient inputs by modeling. Proc. Nutr. Soc. 35:383–391. Whittemore, C. T. 1983. Development of recommended energy and protein allowances for young pigs. Agric. Syst. 11:159–186. Whittemore, C. T., and C. A. Morgan. 1990. Model components for the determination of energy and protein requirements for breeding sows: A review. Livestock Prod. Sci. 26:1–37. Whittemore, C. T., and H. Yang. 1989. Physical and chemical composition of the body of breeding sows with differing body subcutaneous fat depth at parturition, differing nutrition during lactation and differing litter size. Anim. Prod. 48:203–212.

15

Whittemore, C. T., A. G. Taylor, G. M. Hillyer, D. Wilson, and C. Stamataris. 1984. Influence of body fat stores on reproductive performance. Anim. Prod. 38:527 (Abstr.). Williams, I. H., W. H. Close, and D. J. A. Cole. 1985. Strategies for sow nutrition: Predicting the response of pregnant animals to protein and energy intake. Pp. 133–147 in Recent Advances in Animal Nutrition, W. Haresign, and D. J. A. Cole, eds. London: Butterworth. Wu, J. F., and R. C. Ewan. 1979. Utilization of energy of wheat and barley by young swine. J. Anim. Sci. 49:1470–1477. Zivkovic, S., and J. P. Bowland. 1970. Influence of substituting higher fiber ingredients for corn on the digestibility of diets and performance of sows and litters. Can. J. Anim. Sci. 50:177–184.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids

2 Protein generally refers to crude protein, which is defined for mixed feedstuffs as the nitrogen content 2 6.25. This definition is based on the assumption that, on average, the nitrogen content is 16 g of nitrogen/100 g of protein. Proteins are composed of amino acids, and it is actually the amino acids that are the essential nutrients. Therefore, the dietary provision of amino acids in correct amounts and proportions determines the adequacy of a dietary protein concentrate. Supplemental nonprotein nitrogen, such as urea, has not produced beneficial responses in swine that were fed practical diets (Hays et al., 1957; Kornegay et al., 1965; Wehrbein et al., 1970).

A few amino acids do not fit neatly into the essential and nonessential classifications. An example is arginine, which is generally classified as an essential amino acid. Swine can synthesize arginine, and arginine synthesis from glutamine has been detected in pig enterocytes prepared within 1 hour of farrowing (Wu and Knabe, 1995). However, synthesis is not adequate to meet nutrient requirements during the early stages of growth (Southern and Baker, 1983). Consequently, the diets of growing swine must contain a source of arginine. However, during postpubertal growth and pregnancy, swine can synthesize arginine at a rate sufficient to meet most or all of their needs (Easter et al., 1974; Easter and Baker, 1976). Synthesis of arginine is probably insufficient to meet the demands of lactation. Proline is not considered an essential amino acid for swine. Initial research by Ball et al. (1986) suggested that very young pigs (1 to 5 kg) were unable to synthesize proline rapidly enough to meet their requirements, and, as a result, a dietary source of proline must be provided. These conclusions were reached on the basis of changes in the oxidation of an indicator amino acid. However, subsequent research from the same laboratory revealed no differences in growth rate between pigs given a diet with almost no dietary supply of proline and pigs fed a diet with supplemental proline (Murphy, 1992). This finding led the author to conclude that proline is not a dietary essential amino acid for neonatal pigs. Furthermore, Chung and Baker (1993) fed a proline-free diet to 5-kg pigs and also observed no growth response to supplemental proline. There are no reports that other classes of swine (greater than 5 kg) require a dietary source of proline. Cysteine can be synthesized from methionine, and therefore it is classified as nonessential. However, cysteine and its oxidation product cystine can satisfy approximately 50 percent of the need for total sulfur amino acids (methionine ` cystine) (Shelton et al., 1951; Becker et al., 1955; Mitchell et al., 1968; Baker et al., 1969; Roth and Kirchgessner,

ESSENTIAL AND NONESSENTIAL AMINO ACIDS Although there are 20 primary amino acids that occur in proteins, not all of them are essential dietary components. Some amino acids can be synthesized by using carbon skeletons (derived primarily from glucose and other amino acids) and amino groups derived from other amino acids present in excess of the requirement. Amino acids synthesized in this manner are termed nonessential (or dispensable). Amino acids that cannot be synthesized, or cannot be synthesized at a sufficient rate to permit optimal growth or reproduction, are termed essential (or indispensable). Although amino acids in both categories are needed at the physiologic or metabolic level, normal swine diets contain adequate amounts of nonessential amino acids or of amino groups for their synthesis. This seems to be true even for low-protein diets that are supplemented with crystalline amino acids (Brudevoid and Southern, 1994). Thus, most of the emphasis in swine nutrition is on the essential amino acids.

16

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids 1989; Chung and Baker, 1992a), and in this way can reduce the need for methionine. Methionine cannot be synthesized from cystine, and therefore it is essential. Methionine can meet the total need for sulfur amino acids in the absence of cystine. Similarly, phenylalanine can meet the total requirement for phenylalanine and tyrosine (aromatic amino acids) because it can be converted to tyrosine. Tyrosine can satisfy at least 50 percent of the total need for these two amino acids (Robbins and Baker, 1977), but it cannot serve as the sole source, because it cannot be converted to phenylalanine. Glutamine is considered to be a conditionally essential amino acid in some species (Lacey and Wilmore, 1990), because it prevents intestinal atrophy under certain conditions. Wu et al. (1996) recently reported that addition of 1 percent glutamine to a corn–soybean meal diet prevented jejunal atrophy in pigs weaned at 21 days during the first week postweaning and increased feed efficiency during the second week postweaning.

AMINO ACIDS IN DIETS Cereal grains, such as corn, sorghum, barley, or wheat, are the primary ingredients of most swine diets and usually provide 30 to 60 percent of the total amino acid requirements. But other sources of protein, such as soybean meal, must be provided to ensure adequate amounts of, and a proper balance among, the essential amino acids. Supplements of crystalline amino acids may also be used to increase intakes of specific amino acids. The protein levels necessary to provide adequate intakes of essential amino acids will depend on the feedstuffs used. Feedstuffs that contain ‘‘high quality’’ proteins (i.e., they have an amino acid pattern relatively similar to the pig’s needs) or mixtures of feedstuffs in which the amino acid pattern in one complements the pattern in another will meet the essential amino acid requirements at lower dietary protein levels than feedstuffs with a less desirable amino acid pattern. This is important if one of the goals is to minimize nitrogen excretion. Another method of reducing dietary protein levels, and thereby reducing nitrogen excretion, is the judicious use of supplements of crystalline amino acids. The amino acid requirements of growing-finishing swine, expressed in terms of dietary concentration, increase as the energy density of the diet increases. Research data (Chiba et al., 1991a,b) indicate that at higher or lower energy densities than those found in standard grain– soybean meal diets, amino acid requirements (expressed as a percentage of the diet) may need to be adjusted upward or downward, respectively.

17

The procedures used for amino acid analyses may cause variations in published estimates of amino acid requirements. Determined values for tryptophan and sulfur amino acids in dietary ingredients, in particular, vary considerably. Tryptophan analysis is difficult because of the relatively low concentration in most feedstuffs and because tryptophan is partially destroyed during standard acid hydrolysis. Consequently, special precautions are necessary, such as hydrolysis with barium hydroxide, sodium hydroxide, or lithium hydroxide, or protection against oxidation in acid. Methionine and cystine undergo oxidation to multiple derivatives, and controlled oxidation of methionine to methionine sulfone and of cystine to cysteic acid must be carried out with performic acid before acid hydrolysis (Williams, 1994). Ratios Among Amino Acids (Ideal Protein) In determining amino acid requirements, a fundamental concept of this publication is that there is an optimal dietary pattern among essential amino acids that corresponds to the needs of the animal. This optimal dietary pattern is often called ‘‘ideal protein.’’ The basis for ideal protein has been discussed by several authors, including the Agricultural Research Council (1981), Fuller and Wang (1990), Baker and Chung (1992), Cole and Van Lunen (1994), and Baker (1997). The concept of an optimal pattern among amino acids has been applied in previous editions of this publication, particularly the ninth edition (National Research Council, 1988), in which the pattern was explicitly listed in a table. However, in the ninth edition the pattern was developed after an examination of the results of experiments to determine amino acid requirements. In the present edition, the pattern was developed primarily from experiments specifically designed for that purpose. Three ideal proteins are used in this publication, one for maintenance, one for protein accretion, and one for milk synthesis. These three patterns, along with a pattern for body tissue protein, are presented in Table 2-1. The ratios for maintenance were calculated by taking the mean of the maintenance requirements for each amino acid determined at the University of Illinois (Baker et al., 1966a,b; Baker and Allee, 1970) and at the Rowett Research Institute (Fuller et al., 1989) and dividing by the maintenance requirement for lysine. The phenylalanine ` tyrosine requirement determined at the University of Illinois was not considered reliable and was not included in the mean. Arginine is not required for maintenance. The value of 1200 was set to reflect the fact that arginine synthesis can satisfy all the maintenance needs and some of the needs for protein accretion. The maintenance requirement for histidine has not been determined, and so the maintenance ratio was set equal to the ratio for protein accretion.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

18

Nutrient Requirements of Swine TABLE 2-1 Ideal Ratios of Amino Acids to Lysine for Maintenance, Protein Accretion, Milk Synthesis, and Body Tissue Amino Acid

Maintenance a

Protein Accretion b

Milk Synthesis c

Body Tissue d

Lysine Arginine Histidine Isoleucine Leucine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

100 1200 32 75 70 28 123 50 121 151 26 67

100 48 32 54 102 27 55 60 93 60 18 68

100 66 40 55 115 26 45 55 112 58 18 85

100 105 45 50 109 27 45 60 103 58 10 69

a Maintenance ratios were calculated based on the data of Baker et al. (1966a,b), Baker and Allee (1970), and Fuller et al. (1989). The negative value for arginine reflects arginine synthesis in excess of the needs for maintenance. b Accretion ratios were derived by starting with ratios from Fuller et al. (1989) and then adjusting to values that produced blends for maintenance ` accretion that were more consistent with recent empirically determined values (Baker and Chung, 1992; Baker et al., 1993; Hahn and Baker, 1995; Baker, 1997). c Milk protein synthesis ratios were those proposed by Pettigrew (1993) based on a survey of the literature; the value of 73 for valine proposed by Pettigrew was modified to 85. d Body tissue protein ratios were from a survey of the literature (Pettigrew, 1993).

The ratios for protein accretion were derived by starting with the ratios proposed by Fuller et al. (1989). However, these ratios were adjusted to values that produced blends for maintenance and accretion which were more consistent with recent empirically determined values (for a discussion, see Baker and Chung, 1992; Baker et al., 1993; Hahn and Baker, 1995; Baker, 1997). The ratios for milk production were from the review of Pettigrew (1993) except that the value of 73 for valine was modified to 85 (see Chapter 3). The ratios for body tissue protein were also from the review of Pettigrew (1993). Although it is recognized that the amino acid composition of body protein changes as a pig matures (Kyriazakis et al., 1993), a fixed pattern was used.

Bioavailability of Amino Acids In most swine diets, a portion of each amino acid that is present is not biologically available to the animal. This is because most proteins are not fully digested and the amino acids are not fully absorbed, and also because not all absorbed amino acids are metabolically available. Diets vary considerably in the proportions of their amino acids that are biologically available. The amino acids in some proteins such as milk products are almost fully bioavailable, whereas those in other proteins such as certain plant seeds are much less so (Southern, 1991; Lewis and Bayley, 1995). Expressing amino acid requirements in terms of bioavailable requirements is, therefore, desirable. However, it means that to formulate swine diets, the bioavailable amino acid content of the ingredients being considered must be known.

The bioavailability of amino acids in the protein of dietary ingredients has been determined for a wide range of protein sources fed to swine (Tanksley and Knabe, 1984; Sauer and Ozimek, 1986; Southern, 1991; Lewis and Bayley, 1995). The primary method to determine bioavailability has been to measure the proportion of a dietary amino acid that has disappeared from the gut when digesta reach the terminal ileum. Values determined in this manner are termed ‘‘ileal digestibilities’’ rather than bioavailabilities because amino acids are sometimes absorbed in a form that cannot be fully used in metabolism. Furthermore, unless a correction is made for endogenous amino acid losses, the complete terminology is ‘‘apparent ileal digestibilities.’’ In this publication, minimum endogenous losses are accounted for, and both requirements and ingredient contents are expressed in terms of ‘‘true’’ (or standardized) ileal digestible amino acids. When apparent digestibilities are determined, feedstuffs with low protein content are undervalued relative to feedstuffs with high protein content because of the relatively greater contribution of endogenous amino acids. True digestibilities correct for this. In addition, because of the way in which ideal protein patterns were determined, these patterns reflect true ileal digestibility rather than apparent ileal digestibility. In general, ileal digestibility values are similar to values determined by other methods such as growth assays (Green and Kiener, 1989; Kovar et al., 1993; Adeola et al., 1994). For feedstuffs exposed to excess heat treatment, however, ileal digestibilities overestimate bioavailabilities of lysine, threonine, methionine, and tryptophan as determined by growth assays (Batterham, 1992, 1994). Apparent and true

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids ileal digestibility coefficients for various feed ingredients are given in Tables 11-5 and 11-6 in Chapter 11. Amino Acid Isomers In all cases, the requirements listed in this publication refer to the L isomer, the form in which most amino acids occur in plant and animal proteins. When crystalline amino acid supplements are provided, DL-methionine can replace the L form in meeting the need for methionine (Reifsnyder et al., 1984; Chung and Baker, 1992b), although there is evidence that the D form may be used less effectively than the L form by very young pigs (Kim and Bayley, 1983). Estimates of the biological activity of D-tryptophan vary from 60 to 100 percent of that of L-tryptophan for the growing pig (Baker et al., 1971; Arentson and Zimmerman, 1985; Kirchgessner and Roth, 1985; Schutte et al., 1988). The activity of the D form may depend on the proportion of D- and L-tryptophan in the diet and on whether the crystalline amino acid is added as D-tryptophan or as DLtryptophan (the racemic mixture). D-Lysine and D-threonine are not used by any of the animal species that have been tested. The values of the D forms of other essential amino acids for the pig are not known. Commercial feedgrade sources of crystalline amino acids include L-lysinezHCl (98.5 percent pure 4 78.8 percent lysine activity), L-threonine (98.5 percent pure), L-tryptophan (98.5 percent pure), DL -methionine (99 percent pure), and DL-methionine hydroxy analog (a liquid that contains 88 percent methionine hydroxy analog). Research has indicated that on a molar basis DL-methionine and DL-methionine hydroxy analog have the same methionine activity for young pigs (Reifsnyder et al., 1984; Chung and Baker, 1992b). In addition, some amino acids can be purchased together in a mixture (e.g., lysine and tryptophan), and others are available in a liquid form (e.g., lysine).

19

effect on swine performance. In contrast, additions of excessive supplements of crystalline amino acids, such as arginine, leucine, and methionine, can reduce feed intake and growth rate (Oestemer et al., 1973; Henry et al., 1976; Southern and Baker, 1982; Hagemeier et al., 1983; Anderson et al., 1984a,b; Edmonds and Baker, 1987a; Edmonds et al., 1987; Brudevoid and Southern, 1994). Large intakes of individual amino acids can lead to a variety of negative syndromes that have been classified as toxicity, antagonism, and imbalance, depending on the nature of the effect. Antagonisms commonly occur among amino acids that are structurally related. An example is the lysine-arginine antagonism in poultry, in which excessive dietary lysine increases the requirement for arginine. In pigs, however, excess lysine does not seem to increase the arginine requirement (Edmonds and Baker, 1987b). An amino acid imbalance may result when diets are supplemented with one or more amino acids other than the limiting amino acid. A reduction in feed intake is common in most of these situations. Swine usually recover rapidly when the offending amino acid is removed from the diet.

AMINO ACID REQUIREMENTS Starting Pigs A summary of recent research on the amino acid requirements of starting pigs (3 to 20 kg) is included in Table 2-2 and a summary of the lysine requirements from these tabular data is shown in Figure 2-1. Based on these data, the total lysine requirements were set as: 5 kg, 1.45 percent; 10 kg, 1.25 percent; 15 kg, 1.15 percent; and 20 kg, 1.05

1.6 Observed

Requirement

AMINO ACID DEFICIENCIES AND EXCESSES There are few characteristic clinical signs of amino acid deficiencies in swine. The primary sign is usually a reduction in feed intake that is accompanied by increased feed wastage, impaired growth, and general unthriftiness. Swine can tolerate high intakes of protein with few specific ill effects, except occasional mild diarrhea. However, feeding high levels of protein (e.g., in excess of 25 percent protein to growing-finishing pigs) is wasteful, contributes to environmental pollution, and usually results in reduced weight gain and feed efficiency. A corn–soybean meal diet contains quantities of certain amino acids (e.g., arginine, leucine, phenylalanine ` tyrosine) in excess of the levels needed for optimal growth, but these excesses have little

Dietary Lysine (%)

1.4 1.2 1.0 0.8 0.6 0.4 0

20

40

60

80

100

Body Weight (kg)

FIGURE 2-1 Lysine requirements of starting, growing, and finishing pigs in research published since 1985. Each block represents an estimated requirement (total lysine basis) plotted against the mean body weight of the pigs in the experiment (final body weight minus initial body weight divided by 2). The line represents an estimate of the lysine requirement (total lysine basis).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

20

Nutrient Requirements of Swine

TABLE 2-2 Amino Acid and Estimated Requirement (g/d) (%)a Lysine 1.27b 1.40 1.20b

– 5.0 –

Research Findings on Amino Acid Requirements of Growing Swine Since 1985

Weight (kg)

Type of Diet

Response Criteria

Commentsc

References

2–6

Semipurified

Diets contained 0.88 to 1.47%

Leibholz and Parks, 1987

5–11

Corn–soybean meal–dried skim milk Semipurified

Weight gain, feed efficiency, nitrogen retention Weight gain, feed efficiency, plasma urea Weight gain, feed efficiency, nitrogen retention Weight gain, feed efficiency

Diets contained 1.10 to 1.70%

Goodband et al., 1988

Diets contained 0.70 to 1.30%

Leibholz and Parks, 1987

Digestible lysine requirement 1.03%

Martinez and Knabe, 1990

Diets contained 1.10 to 1.50%.

Lepine et al., 1991

Diets contained 1.15, 1.25, and 1.35% Diets contained 0.95, 1.10, and 1.25% 1.30% was superior to 0.70 or 1.00%. Two energy concentrations and two thermal environments 1.48% was superior to 1.24%

Kornegay et al., 1993

Schenck et al., 1992a,b

Boars

Campbell et al., 1988b

Diets contained 0.75 to 1.25%

Thaler et al., 1986

Diets contained 0.80 to 1.25%

Weaver et al., 1988

A supplement of 0.20% lysine improved performance over the basal diet (0.86%) Lysine requirement 1.08 g/MJ of DE Diets contained 1.16 to 1.34%

Hamilton and Veum, 1986

5–16

1.15

7.2

6–10

1.10

4.3

6–11

1.25



7–10

Corn–peanut meal–soybean meal Corn–soybean meal–whey Corn–soybean meal–whey Corn–soybean meal–whey Corn–soybean meal–whey

1.25

7.4

7–17

1.30

11.0

7–25

1.48

9.1

8–19



1.31

13.0

8–20

1.10

7.7

8–20

Wheat–starch–mixed protein supplements Corn–soybean meal

0.95



8–20

Corn–soybean meal–whey Corn–soybean meal

Weight gain, feed efficiency Weight gain, feed efficiency Weight gain, feed efficiency Weight gain, feed efficiency, protein accretion Weight gain, feed efficiency Protein accretion Weight gain, feed efficiency Weight gain, blood urea nitrogen Weight gain, feed efficiency

1.06

7.7

8–21

1.49

13.3

8–25

1.34

9.7

9–19

1.34

14.3

9–26

1.05



10–20

Wheat–soybean meal Wheat–soybean meal Wheat–barley–soybean meal–oat groats Corn–soybean meal

0.98

10.0

10–20

Corn–soybean meal

1.07

14.7

18–45

Barley–soybean meal–fishmeal

Weight gain, feed efficiency

1.20

15.7

20–45

Semipurified

0.94

12.4

20–45

Semipurified

1.08

13.5

20–45

Barley–mixed protein supplements

Weight gain, feed efficiency, carcass traits, protein accretion Weight gain, feed efficiency, carcass traits, protein accretion Weight gain, feed efficiency, protein accretion

0.99

14.0

20–50

Mixed cereal and protein supplements

Weight gain, feed efficiency Weight gain, feed efficiency Weight gain, feed efficiency Weight gain, feed efficiency Weight gain, nitrogen gain, lysine gain

Weight gain, feed efficiency

Lysine requirement 0.95 g/MJ of DE Diets contained 0.95, 1.05, and 1.15% Higher estimates if nitrogen or lysine gains are used as the criterion Supplemental lysine provided by crystalline lysine and by soybean meal Boars. Limit feeding (3.0 2 maintenance). Ileal digestible lysine requirement 0.76–0.82 g/MJ of DE Gilts. Limit feeding (3.0 2 maintenance). Ileal digestible lysine requirement 0.58–0.65 g/MJ of DE Limit feeding (3.0 2 maintenance). Ileal digestible lysine requirement 0.60 g/MJ of DE Humid tropical conditions

Copyright © National Academy of Sciences. All rights reserved.

Mahan et al., 1993

Danielsen et al., 1989

Gatel et al., 1992 ´ ´ Gatel and Fekete, 1989 Nam and Aherne, 1994 Kornegay et al., 1993 Gahl et al., 1994 Fuller et al., 1986 Batterham et al., 1990

Batterham et al., 1990

Bikker et al., 1994

Kuan et al., 1986

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids TABLE 2-2

(continued)

Amino Acid and Estimated Requirement (g/d) (%)a

Weight (kg)

1.04 0.99

17.6 20.0

20–50 20–50

0.76

16.0

20–50

1.09

17.9

20–50

1.03

16.9

20–50

1.02

19.0

1.02

Type of Diet

Response Criteria

Commentsc

References

Lean growth rate Weight gain, feed efficiency

Boars Boars. Diets contained 0.70 to 1.40%

Giles et al., 1986 Giles et al., 1987

Weight gain, feed efficiency

Gilts. Diets contained 0.70 to 1.40%

Giles et al., 1987

Weight gain, feed efficiency

Boars. Lysine requirement 0.75 g/MJ of DE

Campbell et al., 1988a

Weight gain, feed efficiency

Gilts. Lysine requirement 0.71 g/MJ of DE

Campbell et al., 1988a

20–50

Barley–soybean meal Barley–soybean meal and wheat soybean meal Barley–soybean meal and wheat soybean meal Wheat plus mixed protein supplements Wheat plus mixed protein supplements Corn–soybean meal

20–50

Corn–soybean meal

Barrows and gilts. Lysine requirement 3.0 g/Mcal of DE Barrows. Lysine requirement 3.0 g/Mcal of DE

Chiba et al., 1991a

23.0

1.02

22.0

20–60

Corn–soybean meal

17.0

21–50

Corn–soybean meal

0.86

17.2

21–50

0.95

19.5

22–52

Corn–peanut meal–soybean meal Sorghum–soybean meal

Slightly higher requirements for pigs treated with porcine somatotropin A supplement of 0.20% lysine improved performance over the basal diet (0.69%) Digestible lysine requirement 0.71%

Krick et al., 1993

0.89

Weight gain, feed efficiency, plasma urea Weight gain, feed efficiency, protein accretion rate, nitrogen retention Weight gain, feed efficiency, protein accretion rate Weight gain, feed efficiency

0.97

17.6

23–57

1.07

17.4

25–55

1.22

21.5

25–95

1.04



26–30

1.03

17.5

27–35

0.65



30–40

0.85 1.13

– 21.2

32–36 33–55

0.96

22.0

34–72

0.86

19.9

40–85

0.65

21.1

42–101

0.76 1.17

– 25.0

44–49 44–63

Weight gain, feed efficiency

Weight gain, feed efficiency, carcass traits Barley–wheat–soybean Weight gain, feed meal–canola efficiency Barley–soybean Weight gain, feed meal–fishmeal efficiency, carcass traits Wheat–barley–fishmeal– Weight gain, carcass soybean meal traits Corn–soybean meal Weight gain, plasma urea, nitrogen retention Corn–soybean meal Weight gain, plasma urea, Corn plus amino Phenylalanine oxidation acid mix Corn–soybean meal Plasma urea Barley–wheat–soybean Weight gain, feed meal–fishmeal efficiency, carcass traits, nitrogen retention Corn–soybean meal Weight gain, feed efficiency, carcass traits, protein accretion rate Mixed cereal and Weight gain, feed protein efficiency, carcass supplements traits Wheat–peanut Weight gain, feed meal–soybean efficiency, carcass meal traits Corn–soybean meal Plasma urea Barley–wheat–wheat Weight gain, feed gluten–soybean efficiency meal

21

Lawrence et al., 1994

Hamilton and Veum, 1986 Martinez and Knabe, 1990

Barrows and gilts

Owen et al., 1994

Diets contained 0.72 to 0.97%

Bell et al., 1988

Limit feeding. Average of boars, barrows, and gilts

Yen et al., 1986a

Boars and gilts. Limit feeding. Improved strain Barrows

McPhee et al., 1991

Barrows. Diets contained 0.75 to 1.35% Boars

Coma et al., 1995b

Coma et al., 1995a

Lin et al., 1986a

Barrows and gilts Boars. Lysine requirement 0.80 g/MJ of DE

Coma et al., 1995a Rao and McCracken, 1990

Digestible lysine requirement 18 g/d

Friesen et al., 1994a

Humid tropical conditions

Kuan et al., 1986

0.65% was superior to 0.55%

Henry et al., 1992a

Barrows and gilts Barrows. Response up the highest lysine concentration fed

Coma et al., 1995a Susenbeth et al., 1994

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

22

Nutrient Requirements of Swine

TABLE 2-2

(continued)

Amino Acid and Estimated Requirement (g/d) (%)a

Weight (kg)

Type of Diet

Response Criteria

Commentsc

References

0.90

26.5

44–104

Corn–soybean meal

0.90% was superior to 0.70% for medium and high lean growth barrows and gilts

Friesen et al., 1994b

0.60

17.4

47–103

Corn–soybean meal

Barrows

Cromwell et al., 1993

0.87

22.3

47–103

Corn–soybean meal

Gilts

Cromwell et al., 1993

0.70

18.6

49–100

Limit feeding. Barrows

Bourdon and Henry, 1988

0.80

20.0

49–100

Muscle gain

Limit feeding. Gilts

Bourdon and Henry, 1988

0.83 0.82

18.6 20.9

50–85 50–90

Wheat–soybean meal–peanut meal Wheat–soybean meal–peanut meal Barley–soybean meal Barley–soybean meal–fishmeal

Weight gain, feed efficiency, carcass traits, protein accretion rate Weight gain, feed efficiency, carcass traits Weight gain, feed efficiency, carcass traits Muscle gain

Boars Limit feeding. Average of boars barrows, and gilts

Giles et al., 1986 Yen et al., 1986b

0.73

16.2

50–90

Boars. Lysine requirement 0.51 g/MJ of DE

Campbell et al., 1988a

0.63

14.1

50–90

Weight gain, feed efficiency

Gilts. Lysine requirement 0.44 g/MJ of DE

Campbell et al., 1988a

0.70

19.3

50–95

Wheat plus mixed protein supplements Wheat plus mixed protein supplements Corn–soybean meal

Lean growth rate Weight gain, feed efficiency, carcass traits Weight gain, feed efficiency

Weight gain, feed efficiency

Hamilton and Veum, 1986

0.68

23.8

50–95

Corn–soybean meal

A supplement of 0.20% lysine improved performance over the basal diet (0.50%) Barrows. Digestible lysine requirement 0.58%

0.75

21.0

50–95

0.80

22.4

52–78

1.13

27.9

55–88

0.80

26.5

59–105

0.60

16.7

62–108

0.80

25.5

63–99

0.79

22.0

63–100

0.72 0.70

– 22.7

70–74 78–109

0.58

22.9

90–110

0.61

20.4

90–110

0.65

22.7

93–104

Weight gain, feed efficiency, plasma urea, carcass traits Corn–soybean meal Weight gain, feed efficiency, plasma urea, carcass traits Sorghum–soybean Weight gain, feed meal efficiency, carcass traits Barley–wheat–soybean Weight gain, feed meal–fishmeal efficiency, carcass traits, nitrogen retention Corn–soybean meal Weight gain, feed efficiency, carcass traits, protein accretion Corn–sesame meal Weight gain, feed efficiency, plasma urea, carcass traits Corn plus mixed Weight gain, feed protein efficiency, carcass supplements traits, nitrogen retention Barley–wheat–wheat Weight gain, feed gluten–soybean efficiency meal Corn–soybean meal Plasma urea Sorghum–soybean Weight gain, feed meal efficiency, carcass traits Corn–soybean meal Weight gain, feed efficiency, plasma urea, carcass traits Corn–soybean meal Weight gain, feed efficiency, plasma urea, carcass traits Corn–soybean meal Weight gain, plasma urea

Hahn et al., 1995

Gilts. Digestible lysine requirement 0.64%

Hahn et al., 1995

Barrows and gilts

Owen et al., 1994

Boars. Lysine requirement 0.80 g/MJ of DE

Rao and McCracken, 1990

No response to levels . 0.80%

Johnston et al., 1993

Diets contained 0.60 to 1.40%

Goodband et al., 1989

Lysine concentrations greater than 0.80% reduced growth performance

Hansen et al., 1994

Barrows

Susenbeth et al., 1994

Barrows and gilts Barrows and gilts

Coma et al., 1995a Owen et al., 1994

Barrows. Digestible lysine requirement 0.49%

Hahn et al., 1995

Gilts. Digestible lysine requirement 0.52%

Hahn et al., 1995

Barrows. Diets contained 0.45 to 1.05%

Coma et al., 1995b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids TABLE 2-2

(continued)

Amino Acid and Estimated Requirement (g/d) (%)a

Weight (kg)

Type of Diet

Response Criteria

Commentsc

Tryptophan 0.23 0.7

5–10

Semipurified

0.19

1.0

6–16

0.16

1.2

6–22

0.14% was inadequate; 0.23% was adequate Apparent ileal digestible requirement 0.15% Diets contained 0.10 to 0.22%

0.16

1.8

10–20

0.23

2.3

10–35

0.16

2.3

15–40

Corn–fishmeal–corn gluten meal Corn–sunflower meal Corn and mixed protein supplements Corn and mixed protein supplements Semipurified

Weight gain, feed efficiency Weight gain, feed efficiency Weight gain, feed efficiency, serum urea Weight gain, feed efficiency

0.18

2.8

17–38

Corn–soybean meal

0.14

1.9

20–45

Sorghum–meat and bone meal

0.13

2.2

22–50

0.17

3.2

25–60

0.13 0.17

– 4.1

30–45 35–105

0.13

3.5

44–99

0.09

2.8

55–97

0.17b

4.3

60–105

Corn–fishmeal–corn gluten meal Mixed cereals and protein supplements Corn–gelatin Mixed cereals and protein supplements Corn–soybean meal–corn gluten meal Corn–fishmeal–corn gluten meal Corn–peas

Threonine 0.66 1.7

2–5

Semipurified

0.70

3.5

5–15

Sorghum–oat groats–soybean meal

0.54

3.8

5–20

Wheat–peanut meal

0.68

4.0

6–16

0.63

5.7

8–21

0.75

7.1

10–25

0.60

9.4

17–38

Sorghum–peanut meal–soybean meal–whey Corn–sunflower meal Wheat–soybean meal Corn–soybean meal

0.70

8.5

17–50

0.73

9.8

20–40

0.57

14.1

35–105

0.55

15.0

45–105

Corn–hominy feed–meat meal Mixed cereals and protein supplements Mixed cereals and protein supplements Mixed cereals and soy flour

References ` Seve et al., 1991 Burgoon et al., 1992 Borg et al., 1987

Apparent ileal digestible requirement 0.14%

Han et al., 1993

Weight gain, feed efficiency

Diets contained 0.13 to 0.25%

Schutte et al., 1988

Weight gain, feed efficiency Weight gain, feed efficiency

Diets contained 0.11 to 0.18%

Henry et al., 1986

Improved response with addition of 0.04% tryptophan to a diet containing 0.14% Diets contained 0.11 to 0.22%

Russell et al., 1986

Apparent ileal digestible requirement 0.10% Diets contained 0.13 to 0.17%

Burgoon et al., 1992

Weight gain, feed efficiency, carcass traits Weight gain, feed efficiency Weight gain, feed efficiency

Batterham and Watson, 1985

Kiener et al., 1988

Phenylalanine oxidation Weight gain, feed efficiency

Boars Apparent ileal digestible requirement 0.135%

Lin et al., 1986b Lenis et al., 1990

Weight gain, feed efficiency

Improved response with addition of 0.04% tryptophan to a diet containing 0.09% Apparent ileal digestible requirement 0.06% Limit feeding. Small reductions in tryptophan had little effect

Henry et al., 1992b

Weight gain, feed efficiency Weight gain, feed efficiency, carcass traits Weight gain, feed efficiency, N retention Weight gain, feed efficiency, plasma amino acids, plasma urea Weight gain, feed efficiency, N retention Weight gain, feed efficiency Weight gain, feed efficiency, serum urea Weight gain, feed efficiency Weight gain, feed efficiency

Burgoon et al., 1992 ¨ Mohn and Susenbeth, 1994

Threonine may have been highly bioavailable Diets contained 0.53 to 0.83%

Leibholz, 1988

Diets contained 0.49 to 0.77%

Leibholz, 1988

Apparent ileal digestible requirement 0.52%

Saldana et al., 1994

Diets contained 0.50 to 0.78%

Borg et al., 1987 ´ ´ Gatel and Fekete, 1989

Diets contained 0.50 to 0.89%

23

Lewis and Peo, 1986

Weight gain, feed efficiency Weight gain, feed efficiency

Improved response with addition of 0.10% threonine to a diet containing 0.50% Apparent ileal digestible requirement 0.44% Apparent ileal digestible requirement 0.59%

Weight gain, feed efficiency

Apparent ileal digestible requirement 0.42%

Lenis et al., 1990

Weight gain, feed efficiency, carcass traits

Apparent ileal digestible requirement 0.42%

Lenis and van Diepen, 1990

Copyright © National Academy of Sciences. All rights reserved.

Russell et al., 1986 Conway et al., 1990 Schutte et al., 1990

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

24

Nutrient Requirements of Swine

TABLE 2-2

(continued)

Amino Acid and Estimated Requirement (g/d) (%)a

Weight (kg)

0.41

12.0

58–96

0.36

10.4

59–102

Methionine ` cystine 0.58 3.2 5–10

Type of Diet

Response Criteria

Commentsc

References

Sorghum–crystalline amino acids Corn–wheat–corn gluten meal–amino acids

Weight gain, feed efficiency Weight gain, feed efficiency, plasma urea, carcass traits

Apparent ileal digestible requirement 0.28% Higher requirement to minimize plasma urea and maximize carcass traits

Saldana et al., 1994

Semipurified

Weight gain, feed efficiency Weight gain, feed efficiency, plasma urea Weight gain, feed efficiency

Bioavailable methionine requirement 0.255% Methionine requirement 0.41% in the presence of adequate cystine Addition of 0.17% methionine to a diet containing 0.55% did not increase performance Bioavailable methionine requirement 0.255% Diets contained 0.40, 0.50, or 0.60%

0.82

3.4

5–13

0.55

3.6

6–18

0.58

5.7

10–20

Semipurified

0.50

12.8

25–85

Barley–lentils

0.57

9.4

30–60

0.53

12.9

35–105

0.45

12.0

50–80

0.47

10.7

60–90

Mixed cereals and protein supplements Mixed cereals and protein supplements Corn–soybean meal–feather meal Mixed cereals and protein supplements

Valine 0.45

11.3

70

2.8

10

Histidine 0.36

Corn–soybean meal–porcine plasma Corn–soybean meal–sugar

Weight gain, feed efficiency Weight gain, feed efficiency, carcass traits Weight gain, feed efficiency

Zimmerman, 1987

Chung and Baker, 1992c Owen et al., 1995 Lovett et al., 1986 Chung and Baker, 1992c Castell and Cliplef, 1990

Limit feeding

Roth and Kirchgessner, 1987

Weight gain, feed efficiency

Apparent ileal digestible requirement 0.42%

Lenis et al., 1990

Weight gain, feed efficiency Weight gain, feed efficiency

Bioavailable methionine ` cystine requirement 0.40% Limit feeding

Chung et al., 1989

Semipurified

Weight gain, feed efficiency, plasma urea, urea excretion

Apparent ileal digestible requirement 4 0.38%

Lewis and Nishimura, 1995

Semipurified

Weight gain, feed efficiency

Bioavailable requirement 0.31%

Izquierdo et al., 1988

Roth and Kirchgessner, 1987

NOTE: Dashes (–) indicate that no information was available. a Values represent total amino acids on a percentage of the diet as-fed basis. b Values represent total amino acids on a percentage of the diet dry matter basis. c 1 MJ 4 239 kcal.

percent. These estimates are also shown in Figure 2-1. These requirements can be described by the equation: Requirement 4 1.793 1 (0.0873 2 BW) ` (0.00429 2 BW2) 1 (0.000089 2 BW3), R2 4 0.9985

(2-1)

where Requirement 4 lysine requirement (percent of the air-dry diet) and BW 4 body weight (kg). Requirements for other amino acids were calculated from lysine using the ratios established for maintenance and protein accretion on a true ileal digestible basis (Table 2-1), even though there are few empirical data to support these ratios. In general, these requirements for starting pigs are slightly higher than those listed in the previous edition of this publication.

Growing-Finishing Pigs Amino acid requirements of growing-finishing pigs are influenced by their genetic capacity to deposit body protein. The amino acid requirements were calculated by the growth model described in Chapter 3. A summary of recent empirical data on the lysine, tryptophan, threonine, methionine ` cystine, valine, and histidine requirements is included in Table 2-2. Some of the lysine estimates in this table were used to validate the model. Figure 2-1 shows estimates of the lysine requirements (total lysine, percentage basis) from these studies along with an estimate of the lysine requirement at various body weights. In general, these lysine requirements are higher than the estimates listed in the previous edition of this publication. The

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids increase in the lysine requirements is attributed to several factors, among which are improved genetics, health, and other environmental conditions in contemporary pigs. Sows Amino acid requirements of gestating sows are influenced by their requirements for maintenance, protein deposition in maternal proteinaceous tissues, and protein deposition in the products of conception. Amino acid requirements of lactating sows are affected by their needs for maintenance and synthesis of milk protein, adjusted for amino acids that become available from maternal body protein if sows lose weight. Amino acid requirements for sows during gestation and lactation were also developed by computer modeling, as described in Chapter 3. A summary of recent references on the lysine requirements of gestating and lactating sows, some of which were used to validate the models, is shown in Table 2-3. In general, the lysine requirements of pregnant sows are slightly higher, and the lysine requirements of lactating sows are considerably higher, than those listed in the previous edition of this publication. The increased estimates for lactation are supported by the results from a number of studies published since the last edition (Cera et al., 1990; Coffey, 1990; Stahly et al., 1990, 1992; Monegue et al., 1993; Sauber et al., 1994; Knabe et al., 1996). These studies have shown that lactating sows nursing large litters produce more milk (as reflected by increased weaning weights of nursing pigs) and lose less maternal body weight when fed 0.75 to 0.90% dietary lysine (45 to 55 g/day) than when fed the lysine levels cited in the previous edition (0.60% lysine, 35 g/day). Boars There has been little research to determine the amino acid requirements for reproduction in the boar (for a review, see Kemp and Den Hartog, 1989). Inadequate protein intake during development delays sexual maturity and reduces sperm output per ejaculation, but recovery from mild undernutrition (a 12 percent crude protein diet) is fairly rapid (Uzu, 1979). Sexually active boars do not seem to have any unusual amino acid requirements. Early experiments (most of them in Eastern Europe) concerning the effects of lysine and methionine supplements on the reproductive functions of boars indicated that sexually active boars may have a relatively high requirement for sulfur amino acids and perhaps ˇ lysine (Moskutelo, 1970; Netesa and ´ Pashkevich, 1971; Fufaev and ¨ Pashkevich, 1972; Tomme and Loskutnikov, 1972; Huhn et al., 1973, 1974; Poppe et al., 1974a,b,c; Pashkevich, 1974, 1976; Zaripova and Shakirov, 1978). Positive responses to methionine and lysine supplements

25

TABLE 2-3 Lysine Requirements of Gestating and Lactating Sows a Azain, M. J., T. Tomkins, and J. S. Sowinski. 1994. Effect of a protein and energy enriched lactation diet in sow and litter performance: Interaction with supplemental milk replacer. J. Anim. Sci. 72(Suppl. 2):65 (Abstr.). Cera, K. R., L. G. Sterling, and D. Warrington. 1990. Effect of lysine level in lactating diets on sow performance over successive reproduction cycles. J. Anim. Sci. 68(Suppl. 1):365 (Abstr.). Coffey, M. T. 1990. Effect of dietary lysine concentration during lactation on reproductive performance of sows. J. Anim. Sci. 68(Suppl. 1):368 (Abstr.). Coma, J., D. R. Zimmerman, and D. Carrion. 1996. Lysine requirement of the lactating sow determined by using plasma urea nitrogen as a rapid response criterion. J. Anim. Sci. 74:1056–1062. Dourmad, J. Y., M. Etienne, and J. Noblet. 1991. Lysine and other amino acid requirements for lactating sows. J. Anim. Sci. 69(Suppl. 1):366 (Abstr.). Dunn, J. M., and V. C. Speer. 1988. Protein requirement of pregnant gilts. J. Anim. Sci. 66(Suppl. 1):145 (Abstr.). Dunn, J. M., and V. C. Speer. 1989. Minimum nitrogen requirement of pregnant swine. J. Anim. Sci. 68(Suppl. 1):119 (Abstr.). Etienne, M., J. Noblet, J. Y. Dourmad, and H. Fortune. 1989. Study of the lysine requirement of sows during lactation. Journ. Rech. Porcine Fr. 21:101–107. Fernandes, L. C. O., J. H. Britt, and M. T. Coffey. 1990. Effect of frequency of feeding and lysine intake on production and reproduction of primiparous sows. J. Anim. Sci. 68(Suppl. 1):367 (Abstr.). Grandhi, R. R. 1986. Effect of energy and lysine levels on reproductive performance of gilts. Can. J. Anim. Sci. 66:1177(Abstr.). Grandhi, R. R. 1988. Effect of nutritional flushing, supplemental fat and supplemental lysine from puberty to breeding and during early gestation on reproductive performance of gilts. Can. J. Anim. Sci. 68:941–951. Grandhi, R. R. 1992. Effect of feeding supplemental fat or lysine during the postweaning period on the reproductive performance of sows with low or high lactation body weight and fat losses. Can. J. Anim. Sci. 72:679–690. Grandhi, R. R. 1994. Apparent absorption and retention of nutrients during the postweaning period in sows fed supplemental fat or lysine. Can. J. Anim. Sci. 74:123–128. Johnston, L. J., J. E. Pettigrew, and J. W. Rust. 1993. Response of maternal-line sows to dietary protein concentration during lactation. J. Anim. Sci. 71:2151–2156. Jones, D. B., and T. S. Stahly. 1995. Impact of amino acid nutrition during lactation on subsequent reproductive function of sows. J. Anim. Sci. 73(Suppl. 1):85 (Abstr.). Kaji, Y., Y. Hatori, S. Furuya, and T. Ishibashi. 1992a. Lysine requirements of gilts during mid and late pregnancy and mid lactating periods. Anim. Sci. Technol. 63:955–963. Kaji, Y., Y. Hatori, S. Furuya, and T. Ishibashi. 1992b. Lysine requirements of multiparous sows during mid and late pregnancy and mid lactating periods. Anim. Sci. Technol. 63:1175–1181. King, R. H. 1991. Response of pregnant gilts to dietary protein as determined by nitrogen retention. J. Anim. Sci. 69(Suppl. 1):361 (Abstr.). King, R. H., M. S. Toner, H. Dove, C. S. Atwood, and W. G. Brown. 1993. The response of first-litter sows to dietary protein level during lactation. J. Anim. Sci. 71:2457–2463. Knabe, D. A., J. H. Brendemuhl, L. I. Chiba, and C. R. Dove. 1996. Supplemental lysine for sows nursing large litters. J. Anim. Sci. 74:1635–1640. Laurin, J. L., R. D. Goodband, J. L. Nelssen, R. D. Richard, and D. R. Keesecker. 1991. Dietary lysine during lactation affects sow and litter performance. J. Anim. Sci. 69(Suppl. 1):109 (Abstr.). Laurin, J. L., J. L. Nelssen, R. D. Goodband, and M. D. Tokach. 1993. The interrelationships between dietary lysine and litter size on sow and litter performance. J. Anim. Sci. 71(Suppl. 1):65 (Abstr.).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

26

Nutrient Requirements of Swine

TABLE 2-3

(continued)

Monegue, H. J., G. L. Cromwell, R. D. Coffey, S. D. Carter, and M. Cervantes. 1993. Elevated dietary lysine levels for sows nursing large litters. J. Anim. Sci. 71(Suppl. 1):67 (Abstr.). NCR-89. 1995. Effect of room temperature and dietary amino acid concentration on performance of lactating sows. J. Anim. Sci. 73(Suppl. 1):250 (Abstr.). Pinheiro, J. W., H. S. Rostagno, R. Sant’Anna, J. Pereira, and P. M. A. Costa. 1986. Nutritional lysine requirement for lactating sows. Rev. Soc. Bras. Zootec. 15:234–240. Sauber, T. E., T. S. Stahly, R. C. Ewan, and N. H. Williams. 1994. Interactive effects of sow genotype and dietary amino acid intake on lactational performance of sows nursing large litters. J. Anim. Sci. 72(Suppl. 2):66 (Abstr.). Speer, V. C. 1990. Partitioning nitrogen and amino acids for pregnancy and lactation in swine: A review. J. Anim. Sci. 68:553–561. Stahly, T. S., G. L. Cromwell, and H. J. Monegue. 1990. Lactational responses of sows nursing large litters to dietary lysine levels. J. Anim. Sci. 68(Suppl. 1):369 (Abstr.). Stahly, T. S., G. L. Cromwell, and H. J. Monegue. 1992. Milk yield responses of sows nursing large litters. J. Anim. Sci. 70(Suppl. 1):238 (Abstr.). Sterling, L. G., and K. R. Cera. 1990. The effect of dietary lysine level during lactation on milk composition and litter gain efficiency over successive reproduction cycles. J. Anim. Sci. 68(Suppl. 1):365 (Abstr.). Thaler, R. C., R. L. Woerman, and D. B. Britzman. 1992. Effect of lysine level in lactation diets on sow performance and milk composition. J. Anim. Sci. 70(Suppl. 1):238 (Abstr.). Tokach, M. D., R. D. Goodband, J. L. Nelssen, J. L. Laurin, and J. A. Hansen. 1992. The effects of an ideal protein lactation diet on sow and litter performance. J. Anim. Sci. 70(Suppl. 1):69 (Abstr.). Tokach, M. D., J. E. Pettigrew, B. A. Crooker, G. D. Dial, and A. F. Sower. 1992. Quantitative influence of lysine and energy intake on yield of milk components in the primiparous sow. J. Anim. Sci. 70:1864–1872. Touchette, K. J., G. L. Allee, M. D. Newcomb, K. M. Halpin, and R. D. Boyd. 1996. Lysine requirement of the lactating primiparous sow. J. Anim. Sci. 74 (Suppl. 1):63 (Abstr.). Weeden, T. L., J. L. Nelssen, R. C. Thaler, G. E. Fitzner, and R. D. Goodband. 1994. Effect of dietary protein and supplemental soybean oil fed during lactation on sow and litter performance through two parities. Anim. Feed Sci. Technol. 45:211–226. Wilson, M. E., H. Stein, N. L. Trottier, D. D. Hall, R. L. Moser, D. E. Orr, and R. A. Easter. 1996. Effect of lysine intake on reproductive performance in first parity sows. J. Anim. Sci. 74 (Suppl. 1): 63 (Abstr.). a Papers published from 1985 to 1996 and abstracts published in the Journal of Animal Science from 1990 to 1996.

above the NRC (1988) requirements have also been reported by Kim and Moon (1990a,b). In other experiments, however, methionine and lysine supplements have not been beneficial (Ju et al., 1985; Van de Kerk and Willems, 1985). Inadequate protein intakes reduce sperm concentration and total sperm count per ejaculate (Yen and Yu, 1985) as well as libido and semen volume (Louis et al., 1994a). Although minimum protein and amino acid requirements have not been established, a low-protein corn–soybean meal diet (10.6 percent protein, 0.44 percent lysine) fed to provide 7.7 g/day of total lysine was inadequate (Louis et al., 1974b). In this research, a corn–soybean meal diet (15.3 percent protein, 0.83 percent lysine) that provided

360 g/day of protein and 18.1 g/day of total lysine maintained good libido and semen characteristics. Yen and Yu (1985) reported that 280 g/day of protein and 11.6 g/day of total lysine were adequate for boars. Meding and Nielsen (1977) found that there was no increase in sperm production when dietary protein concentration was increased from 15.4 to 18.4 percent. Similarly, Kemp et al. (1988) reported that a diet containing 22.2 percent protein (1.20 percent lysine) did not increase sperm production and semen quality relative to a diet containing 14.5 percent protein (0.68 percent lysine). Because feed intake of adult boars is usually limited to avoid excess weight gain, the daily intakes of amino acids are more important than the dietary amino acid concentrations.

REFERENCES Adeola, O., B. V. Lawrence, and T. R. Cline. 1994. Availability of amino acids for 10- to 20-kilogram pigs: Lysine and threonine in soybean meal. J. Anim. Sci. 72:2061–2067. Agricultural Research Council. 1981. The Nutrient Requirements of Pigs: Technical Review. Rev. ed. Slough, England. Commonwealth Agricultural Bureaux. xxii, 307 pp. Anderson, L. C., A. J. Lewis, E. R. Peo, Jr., and J. D. Crenshaw. 1984a. Effect of various dietary arginine:lysine ratios on performance, carcass composition and plasma amino acid concentrations of growing-finishing swine. J. Anim. Sci. 58:362–368. Anderson, L. C., A. J. Lewis, E. R. Peo, Jr., and J. D. Crenshaw. 1984b. Effect of excess arginine with and without supplemental lysine on performance, plasma amino acid concentrations and nitrogen balance of young swine. J. Anim. Sci. 58:369–377. Arentson, B. E., and D. R. Zimmerman. 1985. Nutritive value of D-tryptophan for the growing pig. J. Anim. Sci. 60:474–479. Baker, D. H. 1997. Ideal amino acid profiles for swine and poultry and their applications in feed formulation. Biokyowa Technical Review— 9. Chesterfield, MO: Nutri-Quest, Inc. Baker, D. H., and G. L. Allee. 1970. Effect of dietary carbohydrate on assessment of the leucine need for maintenance of adult swine. J. Nutr. 100:277–280. Baker, D. H., and T. K. Chung. 1992. Ideal protein for swine and poultry. BioKyowa Technical Review—4. Chesterfield, MO: Nutri-Quest, Inc. Baker, D. H., D. E. Becker, H. W. Norton, A. H. Jensen, and B. G. Harmon. 1966a. Quantitative evaluation of the threonine, isoleucine, valine and phenylalanine needs of adult swine for maintenance. J. Nutr. 88:391–396. Baker, D. H., D. E. Becker, H. W. Norton, A. H. Jensen, and B. G. Harmon. 1966b. Quantitative evaluation of the tryptophan, methionine and lysine needs of adult swine for maintenance. J. Nutr. 89:441–447. Baker, D. H., W. W. Clausing, B. G. Harmon, A. H. Jensen, and D. E. Becker. 1969. Replacement value of cystine for methionine for the young pig. J. Anim. Sci. 29:581–584. Baker, D. H., N. K. Allen, J. Boomgaardt, G. Graber, and H. W. Norton. 1971. Quantitative aspects of D- and L-tryptophan utilization by the young pig. J. Anim. Sci. 33:42–46. Baker, D. H., J. D. Hahn, T. K. Chung, and Y. Han. 1993. Nutrition and Growth: The application of an ideal protein for swine growth. Pp. 133–139 in Growth of the Pig. Wallingford, U.K.: CAB International. Ball, R. O., J. L. Atkinson, and H. S. Bayley. 1986. Proline as an essential amino acid for the young pig. Br. J. Nutr. 55:659–668.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids Batterham, E. S. 1992. Availability and utilization of amino acids for growing pigs. Nutr. Res. Rev. 5:1–18. Batterham, E. S. 1994. Ileal digestibilities of amino acids in feedstuffs for pigs. Pp. 113–131 in Amino Acids in Farm Animal Nutrition. Wallingford, U.K.: CAB International. Batterham, E. S., and C. Watson. 1985. Tryptophan content of feeds, limitations in diets and requirement for growing pigs. Anim. Feed Sci. Technol. 13:171–182. Batterham, E. S., L. M. Andersen, D. R. Baigent, and E. White. 1990. Utilization of ileal digestible amino acids by growing pigs: Effect of dietary lysine concentration on efficiency of lysine retention. Br. J. Nutr. 64:81–94. Becker, D. E., A. H. Jensen, S. W. Terrill, and H. W. Norton. 1955. The methionine-cystine need of the young pig. J. Anim. Sci. 14:1086–1094. Bell, J. M., M. O. Keith, and C. S. Darroch. 1988. Lysine supplementation of grower and finisher pig diets based on high protein barley, wheat and soybean meal or canola meal, with observations on thyroid and zinc status. Can. J. Anim. Sci. 68:931–940. Bikker, P., M. W. A. Verstegen, R. G. Campbell, and B. Kemp. 1994. Digestible lysine requirement of gilts with high genetic potential for lean gain, in relation to the level of energy intake. J. Anim. Sci. 72:1744–1753. Borg, B. S., G. W. Libal, and R. C. Wahlstrom. 1987. Tryphtophan and threonine requirements of young pigs and their effects on serum calcium, phosphorus and zinc concentrations. J. Anim. Sci. 64:1070–1078. Bourdon, D., and Y. Henry. 1988. Lysine requirement of the finishing pig according to sex. Journ. Rech. Porcine Fr. 20:409–414. Brudevoid, A. B., and L. L. Southern. 1994. Low-protein, crystalline amino acid-supplemented, sorghum-soybean meal diets for the 10- to 20-kilogram pig. J. Anim. Sci. 72:638–647. Burgoon, K. G., D. A. Knabe, and E. J. Gregg. 1992. Digestible tryptophan requirements of starting, growing, and finishing pigs. J. Anim. Sci. 70:2493–2500. Campbell, R. G., M. R. Taverner, and D. M. Curic. 1988a. The effects of sex and live weight on the growing pig’s response to dietary protein. Anim. Prod. 46:123–130. Campbell, R. G., M. R. Taverner, and C. J. Rayner. 1988b. The tissue and dietary protein and amino acid requirements of pigs from 8.0 to 20.0 kg live weight. Anim. Prod. 46:283–290. Castell, A. G., and R. L. Cliplef. 1990. Methionine supplementation of barley diets containing lentils (Lens culinaris) or soybean meal: Live performance and carcass responses by gilts fed ad libitum. Can. J. Anim. Sci. 70:329–332. Cera, K. R., L. G. Sterling, and D. Warrington. 1990. Effect of lysine level in lactating diets on sow performance over successive reproduction cycles. J. Anim. Sci. 68(Suppl. 1):365 (Abstr.). Chiba, L. I., A. J. Lewis, and E. R. Peo, Jr. 1991a. Amino acid and energy interrelationships in pigs weighing 20 to 50 kilograms: I. Rate and efficiency of weight gain. J. Anim. Sci. 69:694–707. Chiba, L. I., A. J. Lewis, and E. R. Peo, Jr. 1991b. Amino acid and energy interrelationships in pigs weighing 20 to 50 kilograms: II. Rate and efficiency of protein and fat deposition. J. Anim. Sci. 69:708–718. Chung, T. K., and D. H. Baker. 1992a. Maximal portion of the young pig’s sulfur amino acid requirement that can be furnished by cystine. J. Anim. Sci. 70:1182–1187. Chung, T. K., and D. H. Baker. 1992b. Utilization of methionine isomers and analogs by pigs. Can. J. Anim. Sci. 72:185–188. Chung, T. K., and D. H. Baker. 1992c. Methionine requirement of pigs between 5 and 20 kilograms of body weight. J. Anim. Sci. 70:1857–1863. Chung, T. K., and D. H. Baker. 1993. A note on the dispensability of proline for weanling pigs. Anim. Prod. 56:407–408. Chung, T. K., O. A. Izquierdo, K. Hashimoto, and D. H. Baker. 1989. Methionine requirement of the finishing pig. J. Anim. Sci. 67:2677–2683.

27

Coffey, M. T. 1990. Effect of dietary lysine concentration during lactation on reproductive performance of sows. J. Anim. Sci. 68(Suppl. 1):368 (Abstr.). Cole, D. J. A., and T. A. Van Lunen. 1994. Ideal amino acid patterns. Pp. 99–112 in Amino Acids in Farm Animal Nutrition. J. P. F. D’Mello, ed. Wallingford, U.K.: CAB International. Coma, J., D. Carrion, and D. R. Zimmerman. 1995a. Use of plasma urea nitrogen as a rapid response criterion to determine the lysine requirement of pigs. J. Anim. Sci. 73:472–481. Coma, J., D. R. Zimmerman, and D. Carrion. 1995b. Interactive effects of feed intake and stage of growth on the lysine requirement of pigs. J. Anim. Sci. 73:3369–3375. Conway, D., W. C. Sauer, L. A. den Hartog, and J. Huisman. 1990. Studies on the threonine requirements of growing pigs based on the total, ileal and faecal digestible contents. Livest. Prod. Sci. 25:105–120. Cromwell, G. L., T. R. Cline, J. D. Crenshaw, T. D. Crenshaw, R. C. Ewan, C. R. Hamilton, A. J. Lewis, D. C. Mahan, E. R. Miller, J. E. Pettigrew, L. F. Tribble, and T. L. Veum. 1993. The dietary protein and(or) lysine requirements of barrows and gilts. J. Anim. Sci. 71:1510–1519. Danielsen, V., B. O. Eggum, and H. Jørgensen. 1989. The effect of increased dietary levels of lysine, methionine and threonine on N-retention and growth in weaned pigs. J. Anim. Sci. 67(Suppl. 1):242 (Abstr.). Easter, R. A., and D. H. Baker. 1976. Nitrogen metabolism and reproductive response of gravid swine fed an arginine-free diet during the last 84 days of gestation. J. Nutr. 106:636–641. Easter, R. A., R. S. Katz, and D. H. Baker. 1974. Arginine: A dispensable amino acid for postpubertal growth and pregnancy of swine. J. Anim. Sci. 39:1123–1128. Edmonds, M. S., and D. H. Baker. 1987a. Amino acid excesses for young pigs: Effects of excess methionine, tryptophan, threonine or leucine. J. Anim. Sci. 64:1664–1671. Edmonds, M. S., and D. H. Baker. 1987b. Failure of excess dietary lysine to antagonize arginine in young pigs. J. Nutr. 117:1396–1401. Edmonds, M. S., Gonyou, H. W., and D. H. Baker. 1987. Effect of excess levels of methionine, tryptophan, arginine, lysine or threonine on growth and dietary choice in the pig. J. Anim. Sci. 65:179–185. Friesen, K. G., J. L. Nelssen, R. D. Goodband, M. D. Tokach, J. A. Unruh, D. H. Kropf, and B. J. Kerr. 1994a. Influence of dietary lysine on growth and carcass composition of high-lean-growth gilts fed from 34 to 72 kilograms. J. Anim. Sci. 72:1761–1770. Friesen, K. G., J. L. Nelssen, J. A. Unruh, R. D. Goodband, and M. D. Tokach. 1994b. Effects of the interrelationship between genotype, sex, and dietary lysine on growth performance and carcass composition in finishing pigs fed to either 104 or 127 kilograms. J. Anim. Sci. 72:946–954. Fufaev, I., and A. Pashkevich. 1972. Effect of synthetic lysine on reproductive functions of boars. Svinovodstvo 36(7):32. Fuller, M. F., and T. C. Wang. 1990. Digestible ideal protein—A measure of dietary protein value. Pig News Info. 11:353–357. Fuller, M. F., J. Wood, A. C. Brewer, K. Pennie, and R. MacWilliam. 1986. The responses of growing pigs to dietary lysine, as free lysine hydrochloride or in soya-bean meal, and the influence of food intake. Anim. Prod. 43:477–484. Fuller, M. F., R. McWilliam, T. C. Wang, and L. R. Giles. 1989. The optimum dietary amino acid pattern for growing pigs. 2. Requirements for maintenance and for tissue protein accretion. Br. J. Nutr. 62:255–267. Gahl, M. J., T. D. Crenshaw, and N. J. Benevenga. 1994. Diminishing returns in weight, nitrogen, and lysine gain of pigs fed six levels of lysine from three ´ ´supplemental sources. J. Anim. Sci. 72:3177–3178. Gatel, F., and J. Fekete. 1989. Lysine and threonine balance and requirements for weaned piglets 10–25 kg liveweight fed cereal-based diets. Livest. Prod. Sci. 23:195–206.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

28

Nutrient Requirements of Swine

´ ´ Gatel, F., G. Buron, and J. Fekete. 1992. Total amino acid requirements of weaned piglets 8 to 25 kg live weight given diets based on wheat and soya-bean meal fortified with free amino acids. Anim. Prod. 54:281–287. Giles, L. R., E. S. Batterham, and E. B. Dettmann. 1986. Amino acid and energy interactions in growing pigs. 2. Effects of food intake, sex and live weight on the responses to lysine concentration in barleybased diets. Anim. Prod. 42:133–144. Giles, L. R., E. S. Batterham, E. B. Dettmann, and R. F. Lowe. 1987. Amino acid and energy interactions in growing pigs. 3. Effects of sex and live weight and cereal on the responses to dietary lysine concentration when fed ad libitum or to a restricted food scale on diets based on wheat or barley. Anim. Prod. 45:493–502. Goodband, R. D., R. H. Hines, J. L. Nelssen, and R. C. Thaler. 1988. The effects of dietary lysine level on performance of pigs weaned at two weeks of age. J. Anim. Sci. 66(Suppl. 1):315 (Abstr.). Goodband, R. D., R. H. Hines, J. L. Nelssen, D. H. Kropf, and G. R. Stoner. 1989. The effects of excess dietary lysine additions on growth performance and carcass characteristics of finishing pigs. J. Anim. Sci. 67(Suppl. 1):260 (Abstr.). Green, S., and T. Kiener. 1989. Digestibilities of nitrogen and amino acids in soya-bean, sunflower, meat and rapeseed meals measured with pigs and poultry. Anim. Prod. 48:157–179. Hagemeier, D. L., G. W. Libal, and R. C. Wahlstrom. 1983. Effects of excess arginine on swine growth and plasma amino acid levels. J. Anim. Sci. 57:99–105. Hahn, J. D., and D. H. Baker. 1995. Optimum ratio of lysine to threonine, tryptophan, and sulfur amino acids for finishing swine. J. Anim. Sci. 73:482–489. Hahn, J. D., R. R. Biehl, and D. H. Baker. 1995. Ideal digestible lysine level for early- and late-finishing swine. J. Anim. Sci. 73:773–784. Hamilton, C. R., and T. L. Veum. 1986. Effect of biotin and(or) lysine additions to corn-soybean meal diets on the performance and nutrient balance of growing pigs. J. Anim. Sci. 62:155–162. Han, Y., T. K. Chung, and D. H. Baker. 1993. Tryptophan requirement of pigs in the weight category 10 to 20 kilograms. J. Anim. Sci. 71:139–143. Hansen, J. A., J. L. Nelssen, R. D. Goodband, and J. L. Laurin. 1994. Interactive effects among porcine somatotropin, the beta-adrenergic agonist salbutamol, and dietary lysine on growth performance and nitrogen balance of finishing swine. J. Anim. Sci. 72:1540–1547. Hays, V. W., G. C. Ashton, C. H. Liu, V. C. Speer, and D. V. Catron. 1957. Studies on the utilization of urea by growing swine. J. Anim. Sci. 16:44–54. ´ ´ Henry, Y., P. H. Duee, and A. Rerat. 1976. Isoleucine requirement of the growing pig and leucine-isoleucine interrelationship. J. Anim. Sci. 42:357–364. ´ ´ Henry, Y., P. H. Duee, A. Rerat, and R. Pion. 1986. Determination of tryptophan requirement of growing pigs between 15 and 40 kg live weight. Nutr. Rep. Int. 34:565–573. ` Henry, Y., Y. Colleaux, and B. Seve. 1992a. Effects of dietary level of lysine and of level and source of protein on feed intake, growth performance, and plasma amino acid pattern in the finishing pig. J. Anim. Sci. 70:188–195. ` ´ Henry, Y., B. Seve, Y. Colleaux, and P. Ganier, C. Saligaut, and P. Jego. 1992b. Interactive effects of dietary levels of tryptophan and protein on voluntary feed intake and growth performance in pigs, in relation to plasma free amino acids and hypothalamic serotonin. J. Anim. Sci. ¨70:1873–1887. ¨ Huhn, U., F. Kleemann, I. Konig, and S. Poppe. 1973. Studies on the sexual performance of male pigs fed rations of varying amino acid composition. 1. The influence of amino acid supply and intensity of service on some semen properties of young and old boars. Arch. Tier¨zucht. 16:347–358. ¨ Huhn, U., S. Poppe, B. Numsen, I. Konig, and F. Kleemann. 1974. Investigations on the sexual productivity of male pigs fed varying levels

of amino acids in the ration. 2. Testicle condition, daily semen yield and fertility performances of intensively used boars fed rations supplemented with synthetic amino acids. Arch. Tierzucht. 17:259–268. Izquierdo, O. A., K. J. Wedekind, and D. H. Baker. 1988. Histidine requirement of the young pig. J. Anim. Sci. 66:2886–2892. Johnston, M. E., J. L. Nelssen, R. D. Goodband, D. H. Kropf, R. H. Hines, and B. R. Schricker. 1993. The effects of porcine somatotropin and dietary lysine on growth performance and carcass characteristics of finishing swine fed to 105 or 127 kilograms. J. Anim. Sci. 71:2986–2995. Ju, J. C., S. P. Cheng, and H. T. Yen. 1985. Effect of amino acid supplemented diets on semen characters of boars. J. Chin. Soc. Anim. Sci. 14(1–2):27–35. Kemp, B., and L. A. Den Hartog. 1989. The influence of energy and protein intake on the reproductive performance of the breeding boar: A review. Anim. Reprod. Sci. 20:103–115. Kemp, B., H. J. G. Grooten, L. A. Den Hartog, P. Luiting, and M. W. A. Verstegen. 1988. The effect of a high protein intake on sperm production in boars at two semen collection frequencies. Anim. Reprod. Sci. 17:103–113. Kiener, T., J. Lougnon, and J. C. Jugy. 1988. Contribution to the study of the tryptophan requirement of the growing pig. Journ. Rech. Porcine Fr. 20:401–407. Kim, K. H., and S. J. Moon. 1990a. Effect of lysine levels on semen quality of boars. Korean J. Anim. Sci. 32:767–771. Kim, K. H., and S. J. Moon. 1990b. Effect of methionine levels on semen quality of boars. Korean J. Anim. Sci. 32:800–804. Kim, K. I., and H. S. Bayley. 1983. Amino acid oxidation by young pigs receiving diets with varying levels of sulphur amino acids. Br. J. Nutr. 50:383–390. Kirchgessner, V. M., and F. X. Roth. 1985. Biologische wirksamkeit von ¨ DL-tryptophan bei mastschweinen. Z Tierphysiol. Tierernahr. Futtermittelkd. 54:135–141. Knabe, D. A., J. H. Brendemuhl, L. I. Chiba, and C. R. Dove. 1996. Supplemental lysine for sows nursing large litters. J. Anim. Sci. 74:1635–1640. Kornegay, E. T., E. R. Miller, D. E. Ullrey, B. H. Vincent, and J. A. Hoefer. 1965. Influence of dietary urea on performance, antibody production and hematology of growing swine. J. Anim. Sci. 24:951–954. Kornegay, E. T., M. D. Lindemann, and V. Ravindran. 1993. Effects of dietary lysine levels on performance and immune response of weanling pigs housed at two floor space allowances. J. Anim. Sci. 71:552–556. Kovar, J. L., A. J. Lewis, T. R. Radke, and P. S. Miller. 1993. Bioavailability of threonine in soybean meal for young pigs. J. Anim. Sci. 71:2133–2139. Krick, B. J., R. D. Boyd, K. R. Roneker, D. H. Beermann, D. E. Bauman, D. A. Ross, and D. J. Meisinger. 1993. Porcine somatotropin affects the dietary lysine requirement and net lysine utilization for growing pigs. J. Nutr. 123:1913–1922. Kuan, K. K., T. K. Mak, and D. J. Farrell. 1986. Effect of dietary energy concentration, protein:energy and lysine:energy ratios on growth of pigs in the humid tropics. Aust. J. Exp. Agric. 26:285–289. Kyriazakis, I., G. C. Emmans, and R. McDaniel. 1993. Whole body amino acid composition of the growing pig. J. Sci. Food Agric. 62:29–33. Lacey, J.M., and D. W. Wilmore. 1990. Is glutamine a conditionally essential amino acid? Nutr. Rev. 48:297–309. Lawrence, B. V., O. Adeola, and T. R. Cline. 1994. Nitrogen utilization and lean growth performance of 20- to 50-kilogram pigs fed diets balanced for lysine:energy ratio. J. Anim. Sci. 72:2887–2895. Leibholz, J. 1988. Threonine supplementation of diets for pigs between 7 and 56 days of age. Anim. Prod. 47:475–480. Leibholz, J., and J. R. Parks. 1987. Lysine supplementation of diets for pigs between 7 and 56 days of age. Anim. Prod. 44:421–426. Lenis, N. P., and J. T. M. van Diepen. 1990. Amino acid requirements of pigs. 3. Requirement for apparent digestible threonine of pigs in different stages of growth. Neth. J. Agric. Sci. 38:609–622.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Proteins and Amino Acids Lenis, N. P., J. T. M. van Diepen, and P. W. Goedhart. 1990. Amino acid requirements of pigs. 1. Requirements for methionine ` cystine, threonine and tryptophan of fast-growing boars and gilts, fed ad libitum. Neth. J. Agric. Sci. 38:577–595. Lepine, A. J., D. C. Mahan, and Y. K. Chung. 1991. Growth performance of weanling pigs fed corn-soybean meal diets with or without dried whey at various L-lysine•HCl levels. J. Anim. Sci.. 69:2026–2032. Lewis, A. J., and H. S. Bayley. 1995. Amino acid bioavailability. Pp. 35–65 in Bioavailability of Nutrients for Animals: Amino Acids, Minerals, and Vitamins. Ammerman, C. B., D. H. Baker, and A. J. Lewis, eds. San Diego, CA: Academic Press. Lewis, A.J., and N. Nishimura. 1995. Valine requirement of the finishing pig. J. Anim. Sci. 73:2315–2318. Lewis, A. J., and E. R. Peo, Jr. 1986. Threonine requirement of pigs weighing 5 to 15 kg. J. Anim. Sci. 62:1617–1623. Lin, F. D., T. K. Smith, and H. S. Bayley. 1986a. Influence of dietary lysine concentration on the oxidation of an indicator amino acid by growing boars. J. Anim. Sci. 63:1179–1183. Lin, F. D., T. K. Smith, and H. S. Bayley. 1986b. Tryptophan requirement of growing swine as determined by the oxidation of an indicator amino acid. J. Anim. Sci. 63:467–471. Louis, G. F., A. J. Lewis, W. C. Weldon, P. S. Miller, R. J. Kittok, and W. W. Stroup. 1994a. The effect of protein intake on boar libido, semen characteristics, and plasma hormone concentrations. J. Anim. Sci. 72:2038–2050. Louis, G. F., A. J. Lewis, W. C. Weldon, P. M. Ermer, P. S. Miller, R. J. Kittok, and W. W. Stroup. 1994b. The effect of energy and protein intakes on boar libido, semen characteristics, and plasma hormone concentrations. J. Anim. Sci. 72:2051–2060. Lovett, T. D., M. T. Coffey, R. D. Miles, and G. E. Combs. 1986. Methionine, choline and sulfate interrelationships in the diet of weanling swine. J. Anim. Sci. 63:467–471. Mahan, D. C., R. A. Easter, G. L. Cromwell, E. R. Miller, and T. L. Veum. 1993. Effect of dietary lysine levels formulated by altering the ratio of corn-soybean meal with or without dried whey and L-lysine•HCl in diets for weanling pigs. J. Anim. Sci. 71:1848–1852. Martinez, G. M., and D. A. Knabe. 1990. Digestible lysine requirement of starter and grower pigs. J. Anim. Sci. 68:2748–2755. McPhee, C. P., K. C. Williams, and L. J. Danials. 1991. The effect of selection for rapid lean growth on the dietary lysine and energy requirements of pigs fed to scale. Livest. Prod. Sci. 27:185–198. Meding, A. J. H., and H. E. Nielsen. 1977. Fortskellige proteinnormers indflydelse pa frugtbarheden hos orner, der anvendes til kunstig saerdoverforing. Statens Husdyrbrugsforog, Copenhagen, Denmark. Mitchell, J. R., Jr., D. E. Baker, B. G. Harmon, H. W. Norton, and A. H. Jensen. 1968. Some amino acid needs of the young pig fed a semisynthetic diet. J. Anim. Sci. 27:1322–1326. ¨ Mohn, S., and A. Susenbeth. 1994. Tryptophan requirement of pigs between 60 and 105 kg live weight. J. Anim. Physiol. Anim. Nutr. 72:252–259. Monegue, H. J., G. L. Cromwell, R. D. Coffey, S. D. Carter, and M. Cervantes. 1993. Elevated dietary lysine levels for sows nursing large litters. J. Anim. Sci. 71(Suppl. 1):67 (Abstr.). ˇ Moskutelo, I. I. 1970. Different amounts of lysine for boars. Svinovodstvo 1:27–28. (Cited in Nutr. Abstr. Rev. 1970. 40:1478.) Murphy, J. M. 1992. Effects of Nutrition and Development on Proline Metabolism in the Neonatal Piglet. M.S. Thesis. University of Guelph, Guelph, Ontario, Canada. December 1992. Nam, D. S., and F. X. Aherne. 1994. The effects of lysine:energy ratio on the performance of weanling pigs. J. Anim. Sci. 72:1247–1256. National Research Council. 1988. Nutrient Requirements of Swine. Ninth revised ed. Washington, D.C.: National Academy Press. 93 pp. Netesa, A., and A. Pashkevich. 1971. Effect of level of lysine in the diet on sperm production. Svinovodstvo 10:34. (Cited in Nutr. Abstr. Rev. 1972. 42:769.)

29

Oestemer, G. A., L. E. Hanson, and R. J. Meade. 1973. Leucine-isoleucine interrelationship in the young pig. J. Anim. Sci. 36:674–678. Owen, K. Q., D. A. Knabe, K. G. Burgoon, and E. J. Gregg. 1994. Selfselection of diets and lysine requirements of growing-finishing swine. J. Anim. Sci. 72:554–564. Owen, K. Q., J. L. Nelssen, R. D. Goodband, M. D. Tokach, L. J. Kats, and K. G. Friesen. 1995. Added dietary methionine in starter diets containing spray-dried blood products. J. Anim. Sci. 73:2647–2654. Pashkevich, A. I. 1974. Changes in the reproductive function of boar sires under the effect of different conditions of lysine in rations. Izv Timiryazevsk-S-kh Akad. 4:180–187. Pashkevich, A. I. 1976. Change in the reproductive function of stud boars with different levels of lysine in the ration. Izv Timiryazevsk-S-kh Akad. 2:168–175. Pettigrew, J. E. 1993. Amino acid nutrition of gestating and lactating sows. BioKyowa MO: Nutri-Quest. ¨ Technical Review—5. Chesterfield, ¨ Poppe, S., U. Huhn, F. Kleemann, and I. Konig. 1974a. Studies on the effect of diet on the production of sperms in young boars and boars used for AI. 1. Effect of diet on the sperm production of boars used for AI. Arch. Tierernahr. 24:499–512. ¨ ¨ Poppe, S., U. Huhn, F. Kleemann, and I. Konig. 1974b. Studies on the effect of diet on the production of sperms in young boars and boars used for AI. 2. Influence of nutrition upon the sperm production in young boars. Arch. Tierernahr. 24:551–565. ¨ ¨ Poppe, S., U. Huhn, F. Kleemann, and I. Konig. 1974c. Studies on the effect of diet on the production of sperms in young boars and boars used for AI. 3. Influence of nutrition upon the sperm production and service efficiency of boars used for artificial insemination. Arch. Tierernahr. 24:637–648. Rao, D. S., and K. J. McCracken. 1990. Protein requirements of boars of high genetic potential for lean growth. Anim. Prod. 51:179–187. Reifsnyder, D. H., C. T. Young, and E. E. Jones. 1984. The use of low protein liquid diets to determine the methionine requirement and the efficacy of methionine hydroxy analogue for the three-week-old pig. J. Nutr. 114:1705–1715. Robbins, K. R., and D. H. Baker. 1977. Phenylalanine requirement of the weanling pig and its relationship to tyrosine. J. Anim. Sci. 45:113–118. Roth, F. X., and M. Kirchgessner. 1987. Biological efficiency of dietary methionine or cystine supplementation with growing pigs: A contribution to the requirement for S-containing amino acids. J. Anim. Physiol. Anim. Nutr. 58:267–280. Roth, F. X., and M. Kirchgessner. 1989. Influence of the methionine:cystine relationship in the feed on the performance of growing pigs. J. Anim. Physiol. Anim. Nutr. 61:265–274. Russell, L. E., R. A. Easter, V. Gomez-Rojas, G. L. Cromwell, and T. S. Stahly. 1986. A note on the supplementation of low-protein, maizesoya-bean meal diets with lysine, tryptophan, threonine and methionine for growing pigs. Anim. Prod. 42:291–295. Saldana, C. I., D. A. Knabe, K. Q. Owen, K. G. Burgoon, and E. J. Gregg. 1994. Digestible threonine requirements of starter and finisher pigs. J. Anim. Sci. 72:144–150. Sauber, T. E., T. S. Stahly, R. C. Ewan, and N. H. Williams. 1994. Interactive effects of sow genotype and dietary amino acid intake on lactational performance of sows nursing large litters. J. Anim. Sci. 72(Suppl. 2):66 (Abstr.). Sauer, W. C., and L. Ozimek. 1986. Digestibility of amino acids in swine: Results and their practical applications. A review. Livestock Prod. Sci. 15:367–388. Schenck, B. C., T. S. Stahly, and G. L. Cromwell. 1992a. Interactive effects of thermal environment and dietary lysine and fat levels on rate, efficiency, and composition of growth of weanling pigs. J. Anim. Sci. 70:3791–3802. Schenck, B. C., T. S. Stahly, and G. L. Cromwell. 1992b. Interactive effects of thermal environment and dietary amino acid and fat levels

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

30

Nutrient Requirements of Swine

on rate and efficiency of growth of pigs housed in a conventional nursery. J. Anim. Sci. 70:3803–3811. Schutte, J. B., E. J. Van Weerden, and F. Koch. 1988. Utilization of DLand L-tryptophan in young pigs. Anim. Prod. 46:447–452. Schutte, J. B., M. W. Bosch, N. P. Lenis, J. de Jong, and J. T. M. van Diepen. 1990. Amino acid requirements of pigs. 2. Requirement for apparent digestible threonine of young pigs. Neth. J. Agric. Sci. ` 38:597–607. ¨ ´ Seve, B., M. C. Meunier-Salaun, M. Monnier, Y. Colleaux, and Y. Henry. 1991. Impact of dietary tryptophan and behavioral type on growth performance and plasma amino acids of young pigs. J. Anim. Sci. 69:3679–3688. Shelton, D. C., W. M. Beeson, and E. T. Mertz. 1951. The effect of methionine and cystine on the growth of weanling pigs. J. Anim. Sci. 10:57–64. Southern, L. L. 1991. Digestible Amino Acids and Digestible Amino Acid Requirements for Swine. BioKyowa Technical Review—2. Chesterfield, MO: Nutri-Quest, Inc. Southern, L. L., and D. H. Baker. 1982. Performance and concentration of amino acids in plasma and urine of young pigs fed diets with excesses of either arginine or lysine. J. Anim. Sci. 55:857–866. Southern, L. L., and D. H. Baker. 1983. Arginine requirement of the young pig. J. Anim. Sci. 57:402–412. Stahly, T. S., G. L. Cromwell, and H. J. Monegue. 1990. Lactational responses of sows nursing large litters to dietary lysine levels. J. Anim. Sci. 68(Suppl. 1):369 (Abstr.). Stahly, T. S., G. L. Cromwell, and H. J. Monegue. 1992. Milk yield responses of sows nursing large litters. J. Anim. Sci. 70(Suppl. 1):238 (Abstr.). Susenbeth, A., R. Schneider, and K. H. Menke. 1994. The effect of protein and lysine intake on growth and protein retention in pigs. J. Anim. Physiol. Anim. Nutr. 71:200–207. Tanksley, T. D., Jr., and D. A. Knabe. 1984. Ileal digestibilities of amino acids in pig feeds and their use in formulating diets. Pp. 75–95 in Recent Advances in Animal Nutrition, W. Haresign and D. J. A. Cole, eds. London: Butterworth. Thaler, R. C., G. W. Libal, and R. C. Wahlstrom. 1986. Effect of lysine levels in pig starter diets on performance to 20 kg and on subsequent performance and carcass characteristics. J. Anim. Sci. 63:139–144.

´ Tomme, M. F., and P. L. Loskutnikov. 1972. Effect of methionine level in rations of adult breeding boars on metabolism and semen production. Dokl. Vses Ord. Lenina Akad. Skh. Nauk. 7:23–25. (Cited in Nutr. Abstr. Rev. 1973. 43:261.) Uzu, G. 1979. Influence of protein feeding on the reproductive performance of 30 to 90 kg young boars. Ann. Zootech. 28:431–441. Van de Kerk, P., and C. M. T. Willems. 1985. The influence of crude protein, lysine and methionine ` cystine on the fertility of boars. Z. Tierphysiol. Tierernaehrg. Futtermittelkde 53:43–49. Weaver, E. M., B. S. Borg, G. W. Libal, and R. C. Wahlstrom. 1988. Effect of lysine levels in starter diets on subsequent performance and carcass characteristics of growing-finishing pigs. J. Anim. Sci. 66(Suppl. 1):314 (Abstr.). Wehrbein, G. F., P. E. Vipperman, Jr., E. R. Peo, Jr., and P. J. Cunningham. 1970. Diammonium citrate and diammonium phosphate as sources of dietary nitrogen for growing-finishing swine. J. Anim. Sci. 31:327–332. Williams, A. P. 1994. Recent developments in amino acid analysis. Pp. 11–36 in Amino Acids in Farm Animal Nutrition. Wallingford, U.K.: CAB International. Wu, G., and D. A. Knabe. 1995. Arginine synthesis in enterocytes of neonatal pigs. Am. J. Physiol. 269:R621–R629. Wu, G., S. A. Meier, and D. A. Knabe. 1996. Dietary glutamine supplementation prevents jejunal atrophy in weaned pigs. J. Nutr. 126:2578–2584. Yen, H. T., and I. T. Yu. 1985. Influence of digestible energy and protein feeding on semen characteristics of breeding boars. In: Efficient Animal Production for Asian Welfare. Proc. 3rd AAAP Animal Science Congress, Seoul, South Korea, 2:610–612. Yen, H. T., D. J. A. Cole, and D. Lewis. 1986a. Amino acid requirements of growing pigs. 7. The response of pigs from 25 to 55 kg live weight to dietary ideal protein. Anim. Prod. 43:141–154. Yen, H. T., D. J. A. Cole, and D. Lewis. 1986b. Amino acid requirements of growing pigs. 8. The response of pigs from 50 to 90 kg live weight to dietary ideal protein. Anim. Prod. 43:155–165. Zaripova, L., and Sh. Shakirov. 1978. Fodder lysine and methionine in diets for boars. Svinovodstvo 5:14–15. (Cited in Nutr. Abstr. Rev., Series B 48:644.) Zimmerman, D. R. 1987. Threonine requirement of finishing pigs. J. Anim. Sci. 65(Suppl. 1):130 (Abstr.).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Estimating Energy and Amino Acid Requirements

3 Quantitative nutrient requirements are not the same for all pigs but vary with changes in genetic strain, gender, health, temperature, stocking density, and other factors. This fact was acknowledged in the previous edition of this publication (National Research Council, 1988), but information available at that time was judged inadequate to estimate nutrient requirements for specific situations. This edition uses a mathematical modeling method to help the user derive situation-specific estimates of nutrient requirements. Fortunately, the variations in nutrient requirements are not random or mysterious. They are related in logical patterns to variations in a small number of variables. For growing pigs, these variables are rate of protein accretion, energy intake, and dietary energy density. In this edition, our understanding of those patterns is used to estimate, by use of mathematical models, the different nutrient requirements for different pigs.

4. Transparency. The user must be able to understand how the models work (all equations are provided in Appendix 1) and be able to evaluate the information used to develop them. 5. Anchored to Empirical Data. Where possible, quantitative relationships built into the models are based on measurements at near the whole-animal level rather than on theoretical values. It is recognized that the emphasis on ease of use and simplicity has a cost. More complex models requiring more data inputs by the user might be able to produce more accurate estimates of requirements over a wider range of conditions. The models are not traditional simulation models. They do not predict pig performance and carcass composition from nutrient intake and other information. They do not move through time, predicting the changes in body weight and composition at each time step (e.g., one day). Rather, these models are simply a structured method for developing factorial estimates of nutrient requirements. They estimate the amount of a nutrient used for each major function of the body (e.g., maintenance, protein accretion, milk production) and sum them to estimate a total daily requirement. In the interest of simplicity, the models address only energy and amino acid intake. For growing-finishing pigs, the model estimates only amino acid requirements, presuming that the pigs are allowed ad libitum consumption of feed. Both energy and amino acid requirements are estimated for the gestating and lactating sow. The models estimate the amounts of nutrients needed to support the level of performance currently found in the herd of interest (e.g., lean growth rate of finishing pigs or litter growth rate of lactating sows). The current level of performance is a result of many factors—genetic, nutritional, health, and environmental—and the models do not attempt to identify those factors that limit the current level of performance. In some cases, current performance may

OVERVIEW OF THE MODELS The following five principles guided the development of the models: 1. Ease of Use. Any method of estimating situationspecific nutrient requirements will be more difficult to use than were those in previous editions of this publication. However, the models in this edition were developed to be easy to use by people with varying levels of nutritional expertise and with limited information about the specific situation. 2. Continued Relevance. The models should be flexible enough to adapt to continued changes in genetics and production systems that will occur during the life of this edition (until it is replaced by its successor). Many of these changes cannot be predicted at present. 3. Simplicity. The models should not only be easy to use, but also structurally simple, so they can be understood readily by users.

31

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirements of Swine

be limited by the amounts of amino acids consumed. In those cases, the models would be expected to predict amino acid requirements near the amounts currently provided, but performance may be improved by increasing those levels. Therefore, when the predicted requirements are near, or above, the levels currently fed, it would be prudent to repeat the measurements of performance with higher dietary amino acid levels, and use the new estimate of performance level in estimating requirements with the model.

140 Whole Body Protein Accretion (g/d)

32

130 120 110 100

2

3

Y = (0.47666 + 0.02147X - 0.00023758X + 0.000000713X ) × 127.5

90 80 10

20

30

40

50

60

70

80

90

100

110

120

130

Body Weight (kg)

GROWING-FINISHING PIGS Requirement for Lysine The daily lysine requirement is the sum of the requirements for maintenance and for protein accretion. MAINTENANCE

The daily true ileal digestible lysine requirement for maintenance is assumed to be 36 mg/kg of metabolic body weight (BW 0.75), based on the data of Wang and Fuller (1989). PROTEIN ACCRETION

The daily amount of lysine needed to support protein accretion is the product of two numbers: (1) the daily amount of protein accreted, and (2) the amount of true digestible lysine needed for each gram of protein accreted. These components are estimated separately. Protein Accretion Rate The protein accretion rate is estimated in two steps. First, the potential rate is estimated; then, if necessary, the rate is decreased, to be consistent with the amount of energy consumed. The potential protein accretion rate varies in different situations as well as within a situation as the pig grows. It is necessary to have an equation that describes for a given situation the potential protein accretion rate (g/day) at each stage of growth, an example of which is shown in Figure 3-1. That equation can be obtained in either of two ways, at the option of the user: 1. It can be provided by the user. Information provided by the user can determine both the overall rate of accretion and the shape of the accretion curve. This is the preferred method of obtaining the equation, but the user should use reliable data measured in the situation of interest. It is not useful to provide assumed data. 2. The other way is to use a default equation provided in the model. To do so, the user must provide an estimate of the mean fat-free carcass lean accretion rate over the growing-finishing period. This mean accretion rate can be

FIGURE 3-1 Potential whole body protein accretion rate of pigs of high–medium lean growth rate with a carcass fat-free lean gain averaging 325 g/d from 20 to 120 kg body weight (default equation of the model). The lean growth rate of 325 g/day is converted to a mean whole-body protein accretion rate of 127.5 g/day (325/2.55 4 127.5).

calculated easily from four items of information that should be readily available to most users: ● ● ●

Carcass weight at slaughter; Percent fat-free lean in the carcass at slaughter; Assumed fat-free lean in the carcass at the beginning of the growing period; ● Number of days in the growth period. Detailed instructions in Appendix 2 will help the user calculate the mean fat-free carcass lean accretion rate in grams/day for the growing-finishing period. The model assumes the mean lean accretion rate is measured over the period of 20 to 120 kg body weight. If the beginning or ending weights are different from 20 or 120 kg, the mean lean accretion rate must be adjusted; these adjustments are also provided in Appendix Table 2-1. The default equation describing potential protein accretion rate versus body weight is derived from the mean carcass fat-free lean accretion rate in two steps: 1. First, the mean carcass fat-free lean accretion rate is converted to mean whole-body protein accretion rate. This is a two-stage conversion, from fat-free lean tissue to protein and from carcass to whole body. The conversion factor is taken as 2.55 g of carcass fat-free lean tissue per gram of whole-body protein, a value derived from several reports cited by Susenbeth and Keitel (1988), plus more recent data reported by Bikker et al. (1996a,b). 2. Second, an equation is used to provide estimates of the potential protein accretion rate at any body weight, expressed relative to the overall mean, as follows: Factor 4 0.47666 ` (0.02147 2 BW) 1 (0.00023758 2 BW 2) ` (0.000000713 2 BW 3)

(3-1)

where Factor is the proportion of the overall mean, and BW is the body weight in kg (the overall mean of the factor

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements is 1.0). The potential protein accretion rate at a given body weight is then determined by multiplying the overall mean for protein accretion rate by the factor. Similarly, the lean gain at a given body weight can also be determined by multiplying the overall mean lean growth rate by the factor. This equation was proposed as a compromise between widely varying shapes of protein accretion curves versus body weight. When the performance of the entire model in estimating lysine requirements was subsequently evaluated, this equation proved to be satisfactory and was clearly superior to the others tested, at least as a component of this model. This is the equation shown in graphic form in Figure 3-1. This approach simply moves the regression curves up or down with variation in overall lean accretion rate, keeping the shape of the curves constant, as shown in Figure 3-2. As discussed previously, the user may provide an accretion curve for carcass lean or whole body protein, which is different than the model’s default curve. An aid for developing different curves is shown in the Appendix (Appendix 3). To this point, an equation describing protein accretion rate versus body weight is provided, either by the user or by the default equation within the model using the mean carcass fat-free lean accretion rate provided by the user. Then the user simply enters the weight of the pigs whose requirements are to be estimated. The model calculates from the equation the potential protein accretion rate at the body weight indicated. The model tests whether the energy intake is adequate to support the potential rate of protein accretion. The amount of energy (or feed) consumed in the specific situation and at the body weight of interest is determined in either of two ways: 1. It can be provided by the user. This is the preferred method, but the energy (or feed) intake data should be

derived in the situation of interest. It is not useful to provide assumed or desired rates of intake. 2. A default equation describing DE intake at each body weight is provided in the model. It is patterned after the equation presented by National Research Council (1986), but modified to account for recent empirical data suggesting greater feed intake during the early growth period and slightly decreased feed intake during late finishing. The modified equation produces estimates of total dietary lysine requirements, on a percentage basis, that are in general agreement with recent empirical data summarized in Table 2-2 assuming a mean fat-free carcass lean growth rate of 325 g/day. The equation for a combination of barrows and gilts is: DE intake (kcal/day) 4 1,250 ` (188 2 BW) 1 (1.4 2 BW 2) ` (0.0044 2 BW 3)

(3-2)

This equation is modified for either barrows or gilts by applying the following adjustment, which is added to the DE intake for barrows or subtracted for gilts: Adjustment (kcal/day) 4 DEI 2 (10.083 ` (0.00385 2 BW) 1 (0.0000235 2 BW2))

(3-3)

where DEI is DE intake (kcal/day) calculated from Equation 3-2. Equation 3-3, a modification of a National Research Council (1986) equation, results in a difference of approximately 0.1 percentage point in the total lysine requirement between finishing barrows and gilts (see Chapter 10, Table 10-3), as suggested by the studies of Cromwell et al. (1993) and Hahn et al. (1995). The shape of the DE intake curves for barrows, gilts, and a 1:1 ratio for the two genders is shown in Figure 3-3. There are also adjustments in predicted DE intake for variations in ambient temperature and in stocking density (space/pig), as in the system presented originally by National Research Council (1986). 13000

150

12000

140

Carcass Fat-Free Lean Gain

130

350 g/d 325 g/d 300 g/d

120 110 100 90

Daily DE Intake (kcal)

Whole Body Protein Accretion (g/d)

33

11000 Gender

10000

Barrow Mixed Gilt

9000 8000 7000 6000 5000 4000

80 10

20

30

40

50

60

70

80

90

100 110 120 130

Body Weight (kg)

10

20

30

40

50

60

70

80

90 100 110 120 130

Body Weight (kg)

FIGURE 3-2 Potential whole body protein accretion rates of pigs of medium, high–medium, and high lean growth rates with carcass fat-free lean gains averaging 300, 325, and 350 g/day from 20 to 120 kg body weight (default equation of the model).

FIGURE 3-3 Estimated daily digestible energy (DE) intakes of barrows, gilts, and a 1:1 ratio of barrows to gilts consuming feed on an ad libitum basis from 20 to 120 kg body weight (default equation of the model).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirements of Swine

The incremental amount of protein accretion supported by an increment of 1 Mcal of DE intake above the intercept of 55 percent of maintenance is calculated by the following equation, which is a modification of the equation of Black et al. (1986): Protein Accretion 4 ((17.5 2 e10.0192BW) ` 16.25) 2 (MPAR/125) 2 (1 ` (0.015 2 (20 1 T)))

(3-4)

where protein accretion is for a given day expressed in g/Mcal DE intake above 55 percent of maintenance; BW is body weight in kg; MPAR is the mean protein accretion rate over the range of 20 to 120 kg BW expressed in g/day; and T is the effective ambient temperature in °C. This equation estimates protein accretion only when accretion is limited by energy intake. In many situations, energy intake is not limiting and further increments in energy intake do not change protein accretion. The relationship is shown graphically in Figure 3-4 for several body weights. Equation 3-4 estimates the slopes of the ascending lines in Figure 3-4. The first term in the right side of the equation is the same as in the original equation of Black et al. (1986) but expressed in units consistent with the rest of the model. It changes the slope of the relationship of protein accretion to energy intake, causing the slope to gradually flatten as the pig increases in body weight. Results of studies conducted since the publication of the original equation are inconclusive but can be interpreted to suggest that the slope should be flatter for larger pigs than for smaller ones (Quiniou et al., 1995), in agreement with the new equation. The second term is an adjustment of the slope for differences in mean potential protein accretion rate, causing the slope to be steeper for pigs with a greater potential protein accretion rate. There is no compelling reason to believe that the two factors must always be closely related, but the evidence available to date (Campbell and Taverner, 1988; Quiniou et al., 1995) suggests that such a relationship occurs, at least in some situations. In the absence of further

Whole Body Protein Gain (g/d)

200 150

Pig Weight (kg)

100

5 25 50 75 100 150

50 0

information, it is judged prudent to make an adjustment consistent with the current empirical data. The final term in the equation is an adjustment of the slope for ambient temperature. It is based on the report of Close and Mount (1978), which showed clearly that the slope of protein accretion on energy intake becomes flatter as ambient temperature increases. The model solves the equation to determine the amount of protein accretion that can be supported by the amount of DE consumed. It then compares this value with the potential protein accretion rate defined above and takes the lower value as the amount of protein actually accreted if amino acid intake is adequate. Lysine Required per Gram of Protein Accreted This parameter was deduced from recent experiments reported in the literature that estimated the lysine requirement of pigs between 20 and 120 kg body weight. The data set was restricted to those experiments in which (1) lysine was clearly the first-limiting amino acid, and (2) whole (empty)body protein accretion was measured, or carcass lean accretion was measured from which whole-body protein accretion could be calculated by dividing by 2.55 (see earlier discussion). There were eight requirement estimates from three publications (Batterham et al., 1990; Bikker et al., 1994b; Hahn et al., 1995). For each estimate, the amount (g/day) of true ileal digestible lysine above maintenance consumed at the requirement was divided by the protein accretion rate (g/day) at that level of lysine intake. This approach does not require the assumption that the relationship of protein accretion to lysine intake is linear. The eight values ranged from 0.094 to 0.157, with a mean of 0.122. As a further check on this value, whole-body protein accretion (g/day) was plotted (Figure 3-5) against true ileal 25

True Digestible Lysine Above Maintenance (g/d)

34

20

Y = 0.123X + 0.207 2 R = 0.66

15

10

5

0 -50

0

20

40

60

80

100

120

140

160

180

200

Whole Body Protein Deposition (g/d)

-100 -1

0

1

2

3

4

5

6

7

8

9

10

11

12

Digestible Energy Intake (Mcal/d)

FIGURE 3-4 Relationship of whole body protein gain and digestible energy intake in pigs from 5 to 150 kg body weight.

FIGURE 3-5 Relationship of daily whole body protein deposition and daily intake of true ileal digestible lysine above maintenance. Based on data from 18 experiments and adapted from a summary by Kerr (1993).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements digestible lysine intake above maintenance (g/day), using data from a wider range of experiments (Campbell et al., 1984, 1985, 1988, 1990; Batterham et al., 1990; Chiba et al., 1991; Bikker, 1994b; Friesen et al., 1994; Hahn et al., 1995). The data set included the experiments mentioned in the previous paragraph, and also (1) experiments in which it is not clear that lysine was the first-limiting amino acid, and (2) experiments for which it was necessary to make further assumptions in order to estimate whole-body protein accretion. All treatments except those above the requirement are plotted. The slope of the regression line suggests that an increment of approximately 0.123 g of true ileal digestible lysine was consumed for each additional gram of protein accreted. The agreement of this number with the one in the previous paragraph (0.122) lends confidence. Therefore, the relationship of lysine required above maintenance to whole-body protein accretion rate in the model is as follows: Lysine 4 0.12 2 PD

(3-5)

where Lysine is the daily requirement for true ileal digestible lysine intake above maintenance in grams, and PD is daily protein deposition in the whole-body in grams. This equation can be considered to encompass two relationships. The first is the lysine content of whole-body protein, a value that varies with protein intake (Bikker et al., 1994a) but is usually within the range of 6.5 to 7.5 g lysine/100 g body protein. The second relationship is the marginal efficiency of use of absorbed lysine for deposition in protein. The regression coefficient in the equation (0.12), when considered along with the lysine content of wholebody protein, reflects a marginal lysine efficiency of 54 to 62 percent. In summary, the lysine requirement for protein accretion is determined from the equation above and is added to the maintenance requirement for lysine to obtain the total daily lysine requirement. All lysine values are in grams of true ileal digestible lysine. An example is shown in Figure 3-6. True Digestible Lysine Requirement (g/d)

18 17 16 15 14 13 12 11 10 10

20

30

40

50

60

70

80

90

100

110

120

130

Body Weight (kg)

FIGURE 3-6 Daily lysine requirement (true ileal digestible basis) of pigs with a mean lean growth rate (carcass fat-free basis) of 325 g/day from 20 to 120 kg body weight as estimated by the model using default equations.

35

Note that the whole-body protein accretion rate is a single adjustment for variations in genetic strain, gender, health, stocking density, as well as interactions among these and other factors. In fact, it is not necessary or possible to enter other descriptions of these variables. The user does not provide a qualitative description of the breed, commercial genetic line, or strain of the pigs. Even if such a description were quantitatively meaningful at the time the model was developed, future genetic progress would diminish the accuracy and usefulness of such a description. Similarly, attempts might be made to define health status by describing the production system employed (e.g., all in/all out animal flow, segregated early weaning, multi-site production), but variations in health status within the type of production system and future development of superior systems would limit the value of such a definition for deriving quantitative estimates of whole-body protein accretion. Requirements for Other Amino Acids Requirements for the essential amino acids other than lysine are also considered to consist of separate components for maintenance and protein deposition. Calculations are based upon the ideal protein system in which requirements for each of the other amino acids are expressed relative to the lysine requirement. The model uses two patterns of ideal protein, one for maintenance and one for protein accretion, as described in Chapter 2. The final blend of the two patterns depends on the relative proportion of lysine needed for maintenance and whole-body protein accretion. The patterns are on a true ileal digestible basis. Expression of Amino Acid Requirements The procedures described above produce estimates of amino acid requirements (true ileal digestible basis) expressed in g/day. The daily DE (or ME) intake is either provided by the user or estimated within the model, so amino acid requirements are easily expressed as g/Mcal DE. The user provides the energy density of the diet (Mcal DE/kg), which allows the calculation of the amount of feed consumed (kg/day). Then the amino acid requirements are calculated as a percentage of the diet, on a true ileal digestible basis (Figure 3-7). The percentage requirements of true ileal digestible amino acids are also expressed as percentage requirements of apparent ileal digestible amino acids and percentage requirements of total amino acids by using the equations given in Table 3-1. The equations in Table 3-1 were derived by calculating the percentages of true and apparent ileal digestible amino acids and total amino acids in diets formulated with varying ratios of corn and soybean meal, using the true and apparent ileal digestibility coefficients reported in Chapter 11, Tables 11-5 and 11-6. It is recognized that these conversions only apply

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

True Digestible Lysine Requirement (%)

36

Nutrient Requirements of Swine

1.1 1.0 0.9 0.8 0.7 0.6 0.5 0.4 10

20

30

40

50

60

70

80

90

100

110

120

13

to corn–soybean meal diets, which emphasize the need to formulate on a true ileal digestible basis for diets that contain other ingredients. The total lysine requirements, expressed as a percent of the diet, generated by the model for pigs with average, medium–high, and high lean growth rates (carcass fat-free lean accretion rates of 300, 325, and 350 g/day, respectively) over the range of 20 to 120 kg body weight are shown in Figure 3-8.

Body Weight (kg)

FIGURE 3-7 Dietary lysine requirement (%, true ileal digestible basis) of pigs with a mean lean growth rate (carcass fat-free basis) of 325 g/day from 20 to 120 kg body weight as estimated by the model using default equations.

GESTATING SOWS Nutrient restriction is used to control weight gain in sows (see discussion in Chapter 1). The model for gestating

TABLE 3-1 Equations for Converting Percentages of Amino Acids from a True Ileal Digestible Basis to an Apparent Ileal Digestible Basis, from an Apparent Ileal Digestible Basis to a True Ileal Digestible Basis, and from a True or Apparent Ileal Digestible Basis to a Total Basis in a Corn–Soybean Meal Diet a True to Apparent b a Arginine Histidine Isoleucine Leucine Lysine Methionine Cystine Methionine ` Cystine Phenylalanine Tyrosine Phenylalanine ` Tyrosine Threonine Tryptophan Valine

10.0089 10.0006 10.0097 0.0157 10.0210 0.0021 10.0002 0.0018 10.0089 10.0030 10.0118 10.0150 10.0074 10.0049

Apparent to True c b 0.9602 0.9456 0.9490 0.9389 0.9524 0.9415 0.9084 0.9246 0.9481 0.9463 0.9473 0.9061 0.9130 0.9230

True to Total d a Arginine Histidine Isoleucine Leucine Lysine Methionine Cystine Methionine ` Cystine Phenylalanine Tyrosine Phenylalanine ` Tyrosine Threonine Tryptophan Valine

0.0213 0.0119 0.0070 10.0452 0.0365 0.0024 0.0029 0.0053 10.0051 0.0031 10.0015 0.0191 0.0043 0.0052

a 0.0092 0.0006 0.0103 10.0167 0.0221 10.0022 0.0003 10.0020 0.0093 0.0031 0.0124 0.0165 0.0081 0.0054

b 1.0414 1.0576 1.0537 1.0651 1.0500 1.0621 1.1008 1.0816 1.0548 1.0567 1.0556 1.1036 1.0953 1.0834

Apparent to Total e b 1.0571 1.0884 1.1198 1.1378 1.0973 1.0948 1.1447 1.1205 1.1261 1.1091 1.1186 1.1373 1.1036 1.1337

a 0.0311 0.0126 0.0180 10.0641 0.0607 0.0000 0.0031 0.0031 0.0054 0.0066 0.0124 0.0379 0.0132 0.0113

a Although

linear relationships are indicated, the actual relationships are more complex. true to apparent ileal digestible amino acids: apparent, % 4 a ` (b 2 true, %). c From apparent to true ileal digestible amino acids: true, % 4 a ` (b 2 apparent, %). d From true ileal digestible amino acids to total amino acids: total, % 4 a ` (b 2 true, %). e From apparent ileal digestible amino acids to total amino acids: total, % 4 a ` (b 2 apparent, %). b From

Copyright © National Academy of Sciences. All rights reserved.

b 1.1009 1.1511 1.1800 1.2119 1.1522 1.1628 1.2603 1.2119 1.1877 1.1721 1.1808 1.2551 1.2088 1.2283

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements Dietary Lysine Requirement (%)

1.30 1.20 1.10 Lean Gain

1.00

350 g/d 325 g/d 300 g/d

0.90 0.80 0.70 0.60 0.50 0.40 10

20

30

40

50

60

70

80

90 100 110 120 130

Body Weight (kg)

FIGURE 3-8 Dietary lysine requirements (%) of pigs of medium, high–medium, and high lean growth rates with carcass fat-free lean gains averaging 300, 325, and 350 g/day from 20 to 120 kg body weight as estimated by the model using default equations. The requirements are for total lysine, assuming a corn–soybean meal mixture.

sows approaches the issue of restriction in either of two ways. First, the user can provide the amount of DE consumed daily as input data, along with the sow’s body weight at breeding and the estimated litter size. The model will then calculate the estimated weight gain (and the composition of that gain) and the amount of each amino acid needed to achieve that gain. Second, the user can provide the desired amount of weight gain as input data, along with the sow’s body weight at breeding and the litter size. The model will then calculate the amount of DE and amino acids needed to achieve that desired gain. The two approaches to the calculations are based on the same quantitative relationships, described below.

37

When the user provides the DE intake, maternal weight gain is determined from the amount of energy available, assuming that the daily energy requirement for growth of the products of conception is 35.8 kcal of ME/pig. The maintenance requirement is 106 kcal of ME/kg 0.75. The energy available for maternal gain is the difference between DE intake converted to ME by the factor of 0.96 and the sum of the energy required for maintenance and the products of conception. The energy (ME) for maternal gain (MEG) is converted to weight gain by the following relationship derived from the data of Beyer et al. (1994): Maternal weight gain (g/day) 4 87 ` (0.12171 2 MEG)

(3-7)

The daily weight gain is the sum of the maternal weight gain and the daily weight gain of the products of conception (19.8 g/day times the number of pigs). The total weight gain for gestation can then be calculated and partitioned to fat and lean as noted above. Requirement for Energy The daily energy requirement is the sum of the requirements for maintenance, for protein and fat accretion, and for thermoregulation. Tissue accretion is the sum of that in the maternal body and the products of conception. MAINTENANCE

The daily maintenance requirement of the gestating sow is considered to be 106 kcal ME/kg BW 0.75 (or 110 kcal DE/kg BW 0.75) (National Research Council, 1988).

Composition of Weight Gain Based on the data of Beyer et al. (1994), the products of conception associated with each fetus are assumed to total 2.28 kg in weight and contain 246 g of protein. The remainder of the weight gain of the gestating sow is in the maternal body and includes both lean and adipose tissues. The proportion of the maternal gain that is fat tissue is estimated based on the following equation from the data of Beyer et al. (1994): Fat tissue accretion (kg) 4 19.08 ` (0.638 2 MG)

(3-6)

where MG is maternal weight gain (kg). Note that the regression coefficient (0.638 in Equation 3-6) will likely vary among animals. The choice of this relationship reflects the assumption that when amino acid requirements are met and energy intake is restricted, it is the amount of energy that sets the limit of fat tissue accretion. The amount of lean tissue that is accreted is the difference between fat tissue accretion and total maternal weight gain.

PROTEIN AND FAT ACCRETION

The amounts of protein and fat accreted daily are calculated as described above, and assuming the gestation length is 115 days. The energy cost of protein accretion is assumed to be 10.6 kcal of ME/g and that of fat accretion to be 12.5 kcal of ME/g. PRODUCTS OF CONCEPTION

The daily energy requirement for the products of conception is 35.8 kcal of ME for each fetus. THERMOREGULATION

Additional energy is required when sows are maintained in a cold environment. In the model, an average 24-hour temperature of 20°C is considered as ideal. The model predicts that a sow with an average gestation weight of 200 kg will need approximately 240 additional kcal of ME (250

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

38

Nutrient Requirements of Swine

kcal of DE) per day for each 1°C below 20°C. No adjustment is made for temperatures above 20°C. The total daily requirement for ME is the sum of the requirements for maintenance, for tissue accretion, for the products of conception, and for thermoregulation. The requirement for DE is the requirement for ME/0.96.

Requirement for Lysine MAINTENANCE

The daily maintenance requirement for true ileal digestible lysine is considered to be 36 mg/kg BW 0.75, as for growing pigs.

Expression of Amino Acid Requirements The procedures described above produce estimates of amino acid requirements (true ileal digestible basis) expressed in g/day. The daily DE intake is either provided by the user or estimated within the model, so amino acid requirements are easily expressed as g/Mcal DE. The user provides the energy density of the diet (Mcal DE/kg), which allows the calculation of the amount of feed consumed (kg/day). Then, the amino acid requirements are calculated as a percentage of the diet, on a true ileal digestible basis. The percentage requirements are also expressed on an apparent ileal digestible basis and a total basis (in a corn–soybean meal diet) by means of the equations given in Table 3-1.

PROTEIN ACCRETION

The daily nitrogen retention is calculated as the sum of maternal protein gain divided by 6.25, and the nitrogen accretion in the products of conception. Regression analysis of the data of King and Brown (1993) shows the true digestible requirement above maintenance for gestating sows to be 0.807 g of lysine/g of nitrogen retained, assuming the true digestibility values shown in Table 11-6 for the ingredients used in the experimental diets (wheat, skim milk powder, and soybean meal). If this parameter is expressed as grams of true ileal digestible lysine above maintenance per gram of accreted protein, the value is 0.129 (0.807/6.25 4 0.129). This value is similar to the corresponding value of 0.12 used in the growth model (see Equation 3-5). The good agreement lends confidence in both parameters. It was suggested in an earlier review (Pettigrew, 1993), based largely on requirement estimates of the National Research Council (1988), that threonine was likely the first-limiting amino acid in the diets used by King and Brown (1993). However, more recent calculations of the Pettigrew (1993) data produced estimates that suggest lysine and threonine are virtually co-limiting in these diets. Thus, it is considered appropriate to treat the response as though lysine were limiting. The total daily requirement for lysine is the sum of the requirements for maintenance and for protein accretion.

Requirements for Other Amino Acids Daily requirements for the other essential amino acids are estimated by a method analogous to the one used for growing pigs. There is a set of requirement ratios for maintenance and another set for protein accretion (Chapter 2). The final blend of the two patterns depends on the relative proportion of lysine needed for maintenance and accretion. The patterns are on a true ileal digestible basis.

LACTATING SOWS Estimation of nutrient requirements for lactating sows is complicated by the sow’s propensity to contribute energy and amino acids retrieved from her own body to help support her milk production. Many sows will not consume enough feed to provide fully the enormous amount of nutrients needed for milk production, and therefore they lose weight. The amount of body reserves used for milk production appears to vary widely. Milk production potential also varies widely among sows, which causes large variations in nutrient requirements. Therefore, the user must describe the pertinent situation by providing information on the number of suckling pigs per litter and the average daily body weight gain of the suckling pigs. The model approaches the calculations in either of two ways. First, the user can provide the amount of energy consumed daily as input, along with litter size and rate of growth of the suckling pigs. The model will then calculate the estimated weight gain or loss, as well as the amino acid requirements to meet the target level of milk production. Second, the user can provide the weight gain or loss of the sow during lactation as input, along with data on litter size and growth rate of the suckling pigs. The model will then calculate the amount of DE and amino acids needed. The model is designed to calculate amino acid requirements for observed levels of milk production. It is tentatively assumed, in the absence of convincing data, that these nutrient levels will also maximize subsequent reproductive performance. This assumption, however, requires further testing. The two approaches to the calculations are based on the quantitative relationships described below.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements Requirement for Energy The daily energy requirement is the sum of the requirements for maintenance, milk production, and thermoregulation. MAINTENANCE

The daily maintenance requirement of the lactating sow is considered to be 106 kcal of ME/kg BW 0.75 (or 110 kcal of DE/kg BW 0.75), the same as for the gestating sow. MILK PRODUCTION

The amount of energy transferred from the sow to the suckling litter in milk is estimated by a rearrangement of the equation of Noblet and Etienne (1989): Milk energy 4 (4.92 2 Litter gain) 1 (90 2 Number of pigs)

(3-8)

where Milk energy is expressed in kcal GE/day and Litter gain is in g/day. The amount of dietary ME required to produce this amount of milk energy is calculated by dividing the milk energy by 0.72, assuming that the marginal efficiency of use of ME for milk production is 72 percent (Noblet and Etienne, 1987). THERMOREGULATION

Lactating sows kept in a cold or hot farrowing house will adjust their energy intake accordingly. The model considers an average 24-hour temperature of 20°C as ideal and it predicts that an additional 310 kcal of dietary ME (323 kcal of DE) will be consumed per day by sows for every 1° below 20°C. Similarly, 310 fewer kcal of ME (323 kcal of DE) will be consumed per day by sows for every 1° above 20°C.

39

This relationship is used directly for calculation of energy balance when the user provides body weight change as an input. Each gram of protein retrieved from the sow’s body is assumed to provide 5.6 kcal of GE toward meeting the energy requirement. The amount of protein is divided by 0.23 to estimate the amount of lean tissue mobilized (assuming lean tissue is 23 percent protein). Subtracting the amount of lean tissue mobilized from the total amount of body weight lost gives an estimate of the amount of adipose mobilized. This adipose tissue is considered to be 90 percent fat, and it is assumed that 9.4 kcal of GE per gram of fat mobilized is available to be applied toward the energy requirement. The total energy from mobilized tissue is used with an efficiency of 0.88 to meet the energy demands of lactation. The regression equation given above indicates that marginal weight loss is 9.42 percent protein by weight. Further calculations from this number show that 9.55 percent of the energy in the mobilized tissues is from protein (using the assumptions described in the previous paragraph). This relationship is used in estimating the amount of protein, fat, and body weight lost when the DE intake (provided as an input) is less than the energy demand. Requirement for Lysine The daily requirement for lysine is the sum of the requirements for maintenance and for milk production, with a reduction to account for the use of the sow’s body protein to provide part of the lysine needed for milk production. MAINTENANCE

The daily maintenance requirement for true ileal digestible lysine is considered to be 36 mg/kg BW 0.75, as for growing pigs. MILK PRODUCTION

ENERGY FROM THE SOW’S BODY

The total energy requirement is modified by the energy associated with changes in body weight during lactation. Regression analysis of the data of Beyer et al. (1994) produced the following relationship: Protein gain 4 1.47 ` (0.0942 2 ADG)

(3-9)

where protein gain is in g/day and ADG is the sow’s average daily gain of body weight in grams. Note that both protein gain and ADG are often negative, reflecting weight loss in the lactating sow. The composition of body weight gain or loss in the lactating sow may vary with several factors, including energy and amino acid intake. However, a constant relationship is used in the model for simplicity.

The requirement is taken to be 22 g of apparent ileal digestible lysine/kg of litter weight gain. This factor was derived from a review of the literature patterned after that of Pettigrew (1993). From several empirical estimates of the lysine requirement to maximize milk production, the requirement and the litter growth rate at the requirement were recorded. The original diet formulations were used to calculate the apparent ileal digestible lysine levels. Reports included in the summary were those used by Pettigrew (1993) in his review (Boomgaardt et al., 1972; Lewis and Speer, 1973; O’Grady and Hanrahan, 1975; Chen et al., 1978; Stahly et al., 1990; Johnston et al., 1993) and one more recent one (Monegue et al., 1993). A total of eight requirement estimates were included. The lysine require-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

40

Nutrient Requirements of Swine

ments were regressed on the litter growth rates (Figure 3-9), to produce the following equation: Lysine 4 1 6.39 ` (0.022 2 Litter gain)

(3-10)

where Lysine is the apparent ileal digestible lysine requirement in g/day, and Litter gain is in g/day. The coefficient, 0.022, is the factor introduced above. The requirement estimate is then converted from apparent to true ileal digestible lysine.

FROM THE SOW’S BODY

The intercept in Equation 3-10 suggests that the sows in these studies were contributing 6.39 g of lysine/day from their body tissues to support milk production. To this number is added the amount of lysine lost unavoidably from the body (the maintenance requirement). The maintenance requirement for the sows in these experiments is estimated to be 1.67 g/day, so the total amount of lysine contributed from the sow’s body is estimated to be the sum of these two numbers, 8.06 g/day. There are alternate methods for the use of the information described above in estimating lysine requirements. The first method is to simply add the maintenance requirement to the total amount of lysine needed to support milk production (0.022 g of lysine/g of litter weight gain). This sum is conceptually the amount needed to prevent mobilization of the sow’s body protein for providing amino acids for milk production. The second method is to subtract from that number the 8.06 g/day described above as the amount that the sow will contribute from her body without reducing milk yield. That smaller number is conceptually the amount needed to maximize milk production while accepting protein loss from the sow’s body. The model uses a third, intermediate method. It subtracts only the 6.39 g/day that is the intercept in Equation 3-9. Note that this estimate of body tissue mobilization is completely independent of the estimates of tissue mobilization that

Apparent Ileal Digestible Lysine Intake (g/d)

50 45 40 35 30

were used in estimating energy requirements and/or weight loss.

Y = -6.390 + 0.022X 2 R = 0.80

Requirements for Other Amino Acids The requirements for the other essential amino acids are calculated from the ratios of amino acid requirements for maintenance (same as in the growth model), the ratios of amino acids required for milk production (taken as the ratios in milk [Pettigrew, 1993], with one modification), and the ratios of amino acids contributed by body protein (Pettigrew, 1993). The data reported by Pettigrew (1993) were generated from a survey of the literature. The ratios of amino acids needed for milk production were modified from those offered by Pettigrew (1993) only in the case of valine. There is now evidence at both the whole-animal level (Richert et al., 1996) and the tissue level (Boyd et al., 1995) that the valine requirement of lactating sows is higher than would be predicted from the amount secreted in milk. Therefore, the ratio of valine to lysine for milk production is increased from the value of 0.73 (Pettigrew, 1993) to a value of 0.85. Setting valine at 85 percent of lysine for milk production was based on the assumption that lysine is first-limiting in corn–soybean meal diets containing up to 1.0 percent total lysine. This ratio of 0.85 results in requirement estimations suggesting that lysine and valine are co-limiting in corn–soybean meal diets containing about 1.0 percent lysine, and that valine is first-limiting at higher amino acid concentrations. Expression of Amino Acid Requirements The procedures described above produce estimates of amino acid requirements (true ileal digestible basis) expressed in g/day. The daily DE intake is either provided by the user or estimated within the model, so amino acid requirements are easily expressed as g/Mcal DE. The user provides the energy density of the diet (Mcal DE/kg), which allows the calculation of the amount of feed consumed (kg/day). Then the amino acid requirements are calculated as a percentage of the diet, on a true ileal digestible basis. The percentage requirements are also expressed on an apparent ileal digestible basis and a total basis (in a corn–soybean meal diet) by using the equations given in Table 3-1.

25 20 15

WEANLING PIGS

10 5 0 500

1000

1500

2000

2500

Litter Weight Gain (g/d)

FIGURE 3-9 Relation of litter growth rate to dietary apparent ileal digestible lysine intake by lactating sows.

The growth model does not estimate energy or amino acid requirements for weanling pigs weighing less than 20 kg body weight, because of insufficient information on biological relationships at this early stage of growth. However, a mathematical equation was used to estimate the

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements 2

3

Y = (1,250 + 188X - 1.4X + 0.0044X ) / 3,400

3.0 2.5 2.0 1.5 1.0

2

Y = (-133 + 251X - 0.99X ) / 3,400

0.5

Lysine (%) 4 1.793 1 (0.0873 2 BW) ` (0.00429 2 BW 2) 1 (0.000089 2 BW 3)

0.0

(3-11)

The apparent and true ileal digestible percentages of lysine were calculated by rearranging the equation involving the coefficients in Table 3-1. It is recognized that these coefficients apply to corn–soybean meal mixes and do not take into account other ingredients (milk and/or blood byproducts) that likely will be in diets for young pigs. The percentage true ileal digestible lysine is converted to grams per day and the lysine requirement for protein accretion is calculated by subtracting the lysine requirement for maintenance. The ratios of other amino acids to lysine for maintenance and accretion are used to calculate the true ileal digestible requirement for each of the other amino acids. The total requirements (maintenance plus accretion) are expressed as a percentage of intake and can then be converted to apparent and total by the equations in Table 3-1. The user should be aware that although this method of estimating the other amino acid requirements based on their ratio to lysine seems logical, there is no experimental evidence to support such a method. DE intake is estimated by a modification of the National Research Council (1986) equation for pigs weighing less than 20 kg body weight, as follows: DE intake (kcal/day) 4 1133 ` (251 2 BW) 1 (0.99 2 BW 2) (3-12) Feed intake is then determined by dividing DE intake by the DE concentration of the diet. Figure 3-11 illustrates

0

20

40

60

80

100

120

Body Weight (kg)

FIGURE 3-11 Estimated daily feed intake of pigs from 3 to 20 kg and from 20 to 120 kg body weight based on the default equations for digestible energy intake in the model divided by the digestible energy concentration of the diet (3,400 kcal/kg).

the estimated feed intake of pigs from 3 to 20 kg and from 20 to 120 kg based on the default equations of the model. The daily amino acid requirements (true, apparent, total) are calculated by multiplying the percentage estimates by the daily feed intakes. The equations estimating amino acids do not take into account differences in genetic potential for lean growth rate or differences in health status, both of which likely have a large impact on the requirements of weanling pigs. Also, gender is not considered. Temperature and space per pig, however, can be entered by the user, and they impact the DE intake estimates. The user should be aware that the growth model does not always allow a smooth transition in the amino acid requirements from the end of the starting phase (19.9 kg body weight) to the beginning of the growing phase (20 kg body weight). This is due to the fact that the model estimates amino acid requirements at 20 kg based on the lean growth rate of the pigs, whereas lean growth rate is not taken into account by the model during the starting phase.

MINERAL AND VITAMIN REQUIREMENTS

1.6 Dietary Lysine Requirement (%)

3.5

Daily Feed Intake (kg)

percentage of total dietary lysine required at a given weight between 3 and 20 kg. The regression equation represents the best-fitting line through the following estimated requirements based on empirical data (see Chapter 2): 1.45% lysine at 5 kg, 1.25% lysine at 10 kg, 1.15% lysine at 15 kg, and 1.05% lysine at 20 kg body weight. The equation (shown in Figure 3-10) is:

41

1.5 2

3

Y = 1.793 - 0.0873X + 0.00429 X - 0.000089 X 1.4 1.3 1.2 1.1 1.0 2

4

6

8

10

12

14

16

18

20

22

Body Weight (kg)

FIGURE 3-10 Dietary lysine requirement (%) of pigs from 3 to 20 kg body weight using the default equation of the model (total basis, assuming a corn–soybean meal diet).

Traditional modeling procedures were not used to estimate the requirements for minerals and vitamins. Instead, estimates were made from empirical experiments. Estimates were made on a dietary concentration basis for six weight ranges of pigs (3 to 5, 5 to 10, 10 to 20, 20 to 50, 50 to 80, and 80 to 120 kg body weight) and for gestating and lactating sows. Exponential equations were then used to fit the midpoints of these weight ranges for starting, growing, and finishing pigs, by means of the following equation. Requirement 4 e a`b(ln BW)`c(ln BW)

Copyright © National Academy of Sciences. All rights reserved.

2

(3-13)

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

42

Nutrient Requirements of Swine

Two examples of how the equation gives the requirement for a mineral (calcium) and a vitamin (riboflavin) compared with the estimated requirements for the various weight categories of pigs from 3 to 120 kg body weight are shown in Figures 3-12 and 3-13, respectively. Note that the equation gives a requirement value that intersects the estimated requirement at approximately the midpoint of the body weight range. The individual coefficients for the prediction equations for the minerals and vitamins are shown in Table 3-2. The daily requirements were calculated by multiplying the predicted dietary concentrations by the daily feed intake. Exponential equations were not used to estimate mineral and vitamin requirements for gestating or lactating sows. Daily requirements of minerals and vitamins for sows were calculated by multiplying the estimated dietary concentrations by the daily feed intake.

EVALUATION OF THE MODELS The models were evaluated in three ways (Black, 1995): (1) simulation of experiments reported in the literature, 1.0

Y = ea + b(lnX) + c(lnX)

Dietary Calcium (%)

0.9

2

0.8 Predicted by Model Requirement (Table 10-5)

0.7 0.6 0.5 0.4 0

20

40

60

80

100

120

140

Body Weight (kg)

FIGURE 3-12 Estimated dietary calcium requirement (%) of pigs from 3 to 120 kg body weight using the generalized exponential equation in the model.

Dietary Riboflavin (mg/kg)

4.80

Y = ea + b(lnX) + c(lnX)

4.30

2

3.80 3.30

Predicted by Model Requirement (Table 10-5)

2.80 2.30 1.80 0

20

40

60

80

100

120

Body Weight (kg)

FIGURE 3-13 Estimated dietary riboflavin requirement (mg/ kg) of pigs from 3 to 120 kg body weight using the generalized exponential equation in the model.

and comparison of simulated to measured requirements; (2) subjective evaluation of the response of model predictions to changes in input values (behavioral analysis); (3) tests of the sensitivity of model predictions to changes in selected model parameters. Growth Model Experimental estimates of lysine requirements listed in Table 2-2 were simulated with the model, and the predicted requirements compared to the requirements estimated directly from the experimental data. Inputs to the model included the mean growth rate of carcass fat-free lean tissue and the feed intake recorded in the experiment. Several reports did not provide adequate information to support a reliable simulation, and these were excluded from the process. The default lean accretion curve was used in all cases. Studies included were reports by Rao and McCracken (1990), Friesen et al. (1994), and Coma et al. (1995a,b). There is a systematic error in this approach that causes the model to underestimate the requirement determined experimentally. Most experimental estimates of the lysine requirement under a given set of conditions are conducted over a significant time period, as the pigs grow several kilograms. During the experimental period, the lysine requirement presumably changes. The pigs would be expected to respond to higher dietary lysine concentrations during the early part of the experiment than later, and this early response would be reflected in the requirement estimate. Therefore, the experimentally determined requirement, expressed as percentage of the diet, is appropriate for pigs near the initial weight. However, the feed intake reported for the experiment is usually for the entire period, so it is necessary when using the model to estimate the requirement of pigs at the midpoint of the experiment. This requirement, as percent of the diet, should be lower than the experimental estimate of the requirement at the beginning of the experiment. In order to minimize this bias, experiments that covered more than 25 kg growth were arbitrarily excluded from the evaluation process. However, some bias remains. The results are summarized in Table 3-3. Overall, the model underestimated the requirement by 2.0 g/day. Examination of the difference in three stages of growth showed a mean difference of 10.8 g/day from 20 to 50 kg body weight, an overestimate of 0.1 g/day from 50 to 80 kg, and an underestimate of 4.4 g/day from 80 to 120 kg. On a percentage of the diet basis, the model underestimated the requirements by 0.08 percentage units. The percentage estimates of the model were close to the measured requirements for the two lighter weight groups, but the model estimates were 0.15 percentage points less than the measured requirement for the heaviest weight class.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements

43

TABLE 3-2 Coefficients Used in the Growth Model to Predict Mineral and Vitamin Requirements (percentage or amount/kg of diet) for Pigs of Various Body Weights a Coefficients Minerals Calcium (%) Phosphorus, total (%) Phosphorus, available (%) Sodium (%) Chlorine (%) Magnesium (%) Potassium (%) Copper (mg) Iodine (mg) Iron (mg) Manganese (mg) Selenium (mg) Zinc (mg) Vitamins Vitamin A (IU) Vitamin D3 (IU) Vitamin E (IU) Vitamin K (menadione) (mg) Biotin (mg) Choline (g) Folacin (mg) Niacin, available (mg) Pantothenic acid (mg) Riboflavin (mg) Thiamin (mg) Vitamin B6 (mg) Vitamin B12 (mg)

a

b

c

R2

0.0658 10.2735 10.0557 10.3897 10.3010 — 11.2375 1.8799 — 4.6600 2.0364 10.6910 4.9230

10.1023 10.0262 10.4160 10.7984 10.8317 — 0.0736 0.0097 — 0.0642 10.4508 10.3236 10.1716

10.0185 10.0244 0.0050 0.0815 0.0724 — 10.0412 10.0391 — 10.0597 0.0317 0.0097 10.0134

0.99 0.99 0.99 0.97 0.95 — 0.99 0.99 — 0.99 0.91 0.89 0.96

8.2033 5.6700 3.4095 — — 0.2659 — 3.4970 2.8651 1.7129 — 1.3009 2.9577

10.3548 10.1722 10.5082 — — 10.6297 — 10.3884 10.3171 10.2314 — 10.5088 0.1840

0.0262 0.0042 0.0628 — — 0.0664 — 0.0094 0.0250 0.0005 — 0.0477 10.1092

0.92 0.89 0.83 — — 0.98 — 0.98 0.99 0.99 — 0.93 0.96

2

a Estimated requirements 4 ea`b(lnBW)`c(lnBW) , where BW is body weight. Body weights used in the derivation of the equations represented the midpoints of the weight ranges of 3 to 5, 5 to 10, 10 to 20, 20 to 50, 50 to 80, and 80 to 120 kg. These equations will give values that approximate the mineral and vitamin requirements for pigs of these weight ranges shown in Table 10-5.

TABLE 3-3

Evaluation of the Growth Model a Carcass fat-free lean gain (g/day)

Total dietary lysine Meas req

Pred req

(grams/day)

(% of diet)

Author

Gender

Regimen

Mean BW (kg)

Coma et al. (1995a) Coma et al. (1995a) Rao and McCracken (1990) Friesen et al. (1994) Friesen et al. (1994) Coma et al. (1995a) Coma et al. (1995a) Coma et al. (1995b)

Barrows Barrows Boars

Ad lib Restrict Ad lib

31.3 31.3 44.0

292 292 412

18.3 13.1 21.2

15.4 11.3 23.1

12.9 11.8 1.9

84.2 86.3 109.0

0.97 1.01 1.12

0.81 0.87 1.26

10.16 10.14 0.14

83.5 86.1 112.5

Gilts Gilts Barrows Barrows Gilts

Ad lib Ad lib Ad lib Restrict Ad lib

44.5 63.8 98.3 98.3 105.5

376 376 292 292 292

21.5 22.2 21.9 22.8 19.1

21.2 22.3 17.2 16.8 16.5

10.3 0.1 14.7 16.0 12.6

98.6 100.5 78.5 73.7 86.4

1.28 1.03 0.61 0.85 0.66

1.24 1.06 0.48 0.63 0.57

10.04 0.03 10.13 10.22 10.09

96.9 102.9 78.7 74.1 86.4

12.0

89.6

10.08

90.1

10.8 0.1 14.4

94.5 100.5 79.5

10.05 0.03 10.15

94.8 102.9 79.8

Overall mean Period means 20 to 50 kg BW 50 to 80 kg BW 80 to 120 kg BW

Diff

% Meas

Meas req

Pred req

Diff

% Meas

a Column headings: Meas req 4 measured lysine requirement; Pred req 4 predicted lysine requirement from the model; Diff 4 difference in the measured and predicted requirement; % Meas 4 predicted requirement as a percentage of the measured requirement.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

44

Nutrient Requirements of Swine

TABLE 3-4

Evaluation of the Lactation Model a,b DE Intake Entered

Author

Meas req

Pred req

Weight Change Entered

Diff

% Meas

Pred req

Diff

% Meas

----------------------------------------------------- Total Dietary Lysine (g/day) ----------------------------------------------------Touchette et al. (1996) Stahly et al. (1992) Coma et al. (1996) Sauber, low-lean, (1996) Sauber, high-lean, (1996) King et al. (1993) Knabe et al. (1996)

49.7 47.0 55.3 42.0 51.0 41.0 42.0

38.4 52.2 54.9 46.6 48.1 40.9 41.8

111.3 5.2 10.4 4.6 12.9 10.1 10.2

77.3 111.1 99.3 111.0 94.3 99.8 99.5

42.3 56.8 49.6 45.8 42.2 40.8 37.2

17.4 9.8 15.7 3.8 18.8 10.2 14.8

85.1 120.9 89.7 109.0 82.7 99.5 88.6

Mean

46.9

46.1

10.7

98.9

45.0

11.9

96.5

Touchette et al. (1996) Stahly et al. (1992) Coma et al. (1996) Sauber, low-lean, (1996) Sauber, high-lean, (1996) King et al. (1993) Knabe et al. (1996)

1.28 0.90 0.83 1.15 1.15 1.08 0.75

0.99 1.00 0.82 1.10 0.92 1.08 0.74

10.29 0.10 10.01 10.05 10.23 0.00 10.01

77.3 111.1 98.8 95.7 80.0 100.0 98.7

0.89 0.92 0.89 1.15 1.11 1.08 0.80

10.39 0.02 0.06 0.00 10.04 0.00 0.05

69.5 102.2 107.2 100.0 96.5 100.0 106.7

Mean

1.02

0.95

10.07

94.5

0.98

10.04

97.5

-------------------------------------------------- Total Dietary Lysine (% of diet) ---------------------------------------------------

a The

evaluation is based on the measured and predicted requirements of total dietary lysine. Two conditions were tested—either DE intake was entered in the model or sow lactation weight change was entered. b Column headings: Meas req 4 measured lysine requirement; Pred req 4 predicted lysine requirement from the model; Diff 4 differences in the measured and predicted requirement; % Meas 4 predicted requirement as a percentage of the measured requirement.

Behavioral analysis showed the model to perform qualitatively as expected and consistent with current nutritional concepts. Sensitivity analysis showed the model to be very sensitive to the parameter that relates the lysine requirement to whole-body protein accretion (0.12 g true ileal digestible lysine/g protein accreted). Gestation Model No reports were identified that provided all of the information needed to appropriately test the gestation model by comparison of simulated to measured requirements. Lactation Model Experimental estimates of lysine requirements were simulated with the model, and the predicted requirements compared to the requirements estimated directly from the experimental data. Inputs to the model were DE density, body weight after farrowing, lactation length, number of pigs in the litter, daily pig weight gain, environmental temperature, and either DE intake or sow weight change during lactation. A total of seven requirement estimates from six reports (Stahly et al., 1992; King et al., 1993; Knabe et al., 1996; Coma et al., 1996; Sauber, 1996; Touchette et al., 1996) were simulated, including studies with both highlean and low-lean genotypes by Sauber (1996). In several of these experiments, performance improved as the dietary lysine level increased all the way to the highest level. In

those cases, the measured requirement was taken to be the highest level fed, even though the requirement for maximum performance may have been higher. This approach is appropriate in evaluation of this model because the model estimates the amount of lysine needed to reach the level of performance attained in the experiment. The results are summarized in Table 3-4. When DE intake was provided as an input, the predicted daily lysine requirement averaged 46.1 g, which is 0.7 g less than the measured requirement (range of 111.3 to 5.2 g). The average of the predicted requirements expressed as percentage of the diet was also slightly less than the average measured requirement, and individual cases ranged from an underestimate of 0.29 percent of the diet to an overestimate of 0.10 percent. When sow weight change during lactation was provided as an input rather than DE intake, the predicted daily requirement averaged 1.9 g less than the measured requirement (range of 18.8 to 9.8 g). When the requirement was expressed as percentage of the diet, the model underestimated the requirement by 0.04 percent relative to the measured values, with a range of 10.39 to 0.06 percent.

REFERENCES Batterham, E. S., L. M. Andersen, D. R. Baigent, and E. White. 1990. Utilization of ileal digestible amino acids by growing pigs: Effect of

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Models for Energy and Amino Acid Requirements dietary lysine concentration on efficiency of lysine retention. Br. J. Nutr. 64:81–94. Beyer, M., W. Jentsch, L. Hoffmann, R. Schiemann, and M. Klein. 1994. Untersuchungen zum energie- and stickstoffumsatz von graviden und lacktierend suan sowievpm saigferkeln 4. Mittielung—Chemische Zusammensetzung and energiegehalt der Konzeptionsprodukte, der reproduktiven Organe und der Lebenmassezunahmmen order -abnahmen bei graviden and laktierenden Sauen. Arch. Anim. Nutr. 46: 7–37. Bikker, P., M. W. A. Verstegen, and M. W. Bosch. 1994a. Amino acid composition of growing pigs is affected by protein and energy intake. J. Nutr. 124:1961–1969. Bikker, P., M. W. A. Verstegen, R. G. Campbell, and B. Kemp. 1994b. Digestible lysine requirement of gilts with high genetic potential for lean gain, in relation to the level of energy intake. J. Anim. Sci. 72:1744–1753. Bikker, P., M. W. A. Verstegen, B. Kemp, and M. W. Bosch. 1996a. Performance and body composition of finishing gilts (45 to 85 kilograms) as affected by energy intake and nutrition in earlier life: I. Growth of the body and body components. J. Anim. Sci. 74: 806–816. Bikker, P., M. W. A. Verstegen, and R. G. Campbell. 1996b. Performance and body composition of finishing gilts (45 to 85 kilograms) as affected by energy intake and nutrition in earlier life: II. Protein and lipid accretion in body components. J. Anim. Sci. 74: 817–826. Black, J. L. 1995. The testing and evaluation of models. Pp. 23–31 in Modeling Growth in the Pig. P. J. Mougham, M. W. A. Verstegen, and M. I. Visser-Reyneveld, eds. EAAP Publication No. 78. Wageningen Pers, Wageningen. Black, J. L., R. G. Campbell, I. H. Williams, K. J. James, and G. T. Davies. 1986. Simulation of energy and amino acid utilization in the pig. Res. Dev. Agric. 3:121–145. Boomgaardt, J., D. H. Baker, A. H. Jensen, and B. G. Harmon. 1972. Effect of dietary lysine levels on 21-day lactation performance of firstlitter sows. J. Anim. Sci. 34:408–410. Boyd, R. D., R. S. Kensinger, R. J. Harrell, and D. E. Bauman. 1995. Nutrient uptake and endocrine regulation of milk synthesis by mammary tissue of lactating sows. J. Anim. Sci. 73 (Suppl. 2):36–54. Campbell, R. G., and M. R. Taverner. 1988. Genotype and sex effects on the relationship between energy intake and protein deposition in growing pigs. J. Anim. Sci. 66:676–686. Campbell, R. G., M. R. Taverner, and D. M. Curic. 1984. Effect of feeding level and dietary protein content on the growth, body composition and rate of protein deposition in pigs growing from 45 to 90 kg. Anim. Prod. 38:233–240. Campbell, R. G., M. R. Taverner, and D. M. Curic. 1985. The influence of feeding level on the protein requirement of pigs between 20 and 45 kg live weight. Anim. Prod. 40:489–496. Campbell, R. G., M. R. Taverner, and C. J. Raynor. 1988. The tissue and dietary protein and amino acid requirements of pigs from 8.0 to 20.0 kg live weight. Anim. Prod. 46:283–290. Campbell, R. G., R. J. Johnson, R. H. King, M. R. Taverner, and D. J. Meisinger. 1990. Interaction of dietary protein content and exogenous porcine growth hormone administration on protein and lipid accretion rates in growing pigs. J. Anim. Sci. 68:3217–3225. Chen, S. Y., J. P. F. D’Mello, F. W. H. Elsley, and A. G. Taylor. 1978. Effect of dietary lysine levels on performance, nitrogen metabolism and plasma amino acid concentrations of lactating sows. Anim. Prod. 27:331–344. Chiba, L. I., A. J. Lewis, and E. R. Peo, Jr. 1991. Amino acid and energy interrelationships in pigs weighing 20 to 50 kilograms: II. Rate and efficiency of protein and fat deposition. J. Anim. Sci. 69:708–718. Close, W. H., and L. E. Mount. 1978. The effects of plane of nutrition and environmental temperature on the energy metabolism of the grow-

45

ing pig. 2. Growth rate, including protein and fat deposition. Br. J. Nutr. 40:423–431. Coma, J., D. R. Zimmerman, and D. Carrion. 1995a. Interactive effects of feed intake and stage of growth on the lysine requirement of pigs. J. Anim. Sci. 73:3369–3375. Coma, J., D. R. Zimmerman, and D. Carrion. 1995b. Relationship of lean tissue growth and other factors to concentration of urea in plasma of pigs. J. Anim. Sci. 73:3649–3656. Coma, J., D. R. Zimmerman, and D. Carrion. 1996. Lysine requirement of the lactating sow determined by using plasma urea nitrogen as a rapid response criterion. J. Anim. Sci. 74:1056–1062. Cromwell, G. L., T. R. Cline, J. D. Crenshaw, T. D. Crenshaw, R. C. Ewan, C. R. Hamilton, A. J. Lewis, D. C., Mahan, E. R. Miller, J. E. Pettigrew, L. F. Tribble, and T. L. Veum. 1993. The dietary protein and (or) lysine requirements of barrows and gilts. J. Anim. Sci. 71:1510–1519. Friesen, K. G., J. L. Nelssen, R. D. Goodband, M. D. Tokach, J. A. Unruh, D. H. Kropf, and B. J. Kerr. 1994. Influence of dietary lysine on growth and carcass composition of high-lean growth gilts fed from 34 to 72 kilograms. J. Anim. Sci. 72:1761–1770. Hahn, J. D., R. R. Biehl, and D. H. Baker. 1995. Ideal digestible lysine level for early- and late-finishing swine. J. Anim. Sci. 73:773–784. Johnston, L. J., J. E. Pettigrew, and J. W. Rust. 1993. Response of maternal-line sows to dietary protein concentration during lactation. J. Anim. Sci. 71:2151–2156. Kerr, B. J. 1993. Optimizing lean tissue deposition in swine. BioKyowa Technical Review—6. Chesterfield, MO: Nutri-Quest. King, R. H., and W. G. Brown. 1993. Interrelationships between dietary protein level, energy intake, and nitrogen retention in pregnant gilts. J. Anim. Sci. 71:2450–2456. King, R. H., M. S. Toner, H. Dove, C. S. Atwood, and W. G. Brown. 1993. The response of first-litter sows to dietary protein level during lactation. J. Anim. Sci. 71:2457–2463. Knabe, D. A., J. H. Brendemuhl, L. J. Chiba, and C. R. Dove. 1996. Supplemental lysine for sows nursing large litters. J. Anim. Sci. 74:1635–1640. Lewis, A. J., and V. C. Speer. 1973. Lysine requirement of the lactating sow. J. Anim. Sci. 37:104–110. Monegue, H. J., G. L. Cromwell, R. D. Coffey, S. D. Carter, and M. Cervantes. 1993. Elevated dietary lysine levels for sows nursing large litters. J. Anim. Sci. 71(Suppl. 1):67 (Abstr.). National Research Council. 1986. Predicting Feed Intake of FoodProducing Animals. Washington, D.C.: National Academy Press. 85 pp. National Research Council. 1988. Nutrient Requirements of Swine (9th Ed.). Washington, D.C.: National Academy Press. 93 pp. Noblet, J., and M. Etienne. 1987. Metabolic utilization of energy and maintenance requirements in lactating sows. J. Anim. Sci. 64: 774–781. Noblet, J., and M. Etienne. 1989. Estimation of sow milk nutrient output. J. Anim. Sci. 67:3352–3359. O’Grady, J. F., and T. J. Hanrahan. 1975. Influence of protein level and amino-acid supplementation of diets fed in lactation on the performance of sows and their litters. Ir. J. Agric. Res. 14:127–135. Pettigrew, J. E. 1993. Amino Acid Nutrition of Gestating and Lactating Sows. BioKyowa Technical Review—5. Chesterfield, MO: NutriQuest. Quiniou, N., J. Noblet, J. van Milgen, and J.-Y. Dourmad. 1995. Effect of energy intake on performance, nutrient and tissue gain, and protein and energy utilization in growing boars. Anim. Sci. 61:133–143. Rao, D. S., and K. J. McCracken. 1990. Protein requirements of boars of high genetic potential for lean growth. Anim. Prod. 51:179–187. Richert, B. T., M. D. Tokach, R. D. Goodband, J. L. Nelssen, J. E. Pettigrew, R. D. Walker, and L. J. Johnston. 1996. Valine requirement of the high-producing lactating sow. J. Anim. Sci. 74:1307–1313.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

46

Nutrient Requirements of Swine

Sauber, T. E. 1996. Impact of Lean Growth Genotype, Dietary Amino Acid Regimen, and Level of Chronic Immune System Activation on Sow Lactational Performance. Ph.D. Dissertation. Iowa State University, Ames. Stahly, T. S., G. L. Cromwell, and H. J. Monegue. 1990. Lactational responses of sows nursing large litters to dietary lysine levels. J. Anim. Sci. 68(Suppl. 1):369 (Abstr.). Stahly, T. S., G. L. Cromwell, and H. J. Monegue. 1992. Milk yield responses of sows nursing large litters. J. Anim. Sci. 70 (Suppl. 1):238 (Abstr.).

Susenbeth, A., and K. Keitel. 1988. Partition of whole body protein in different body fractions and some constants in body composition in pigs. Livestock Prod. Sci. 20:37–52. Touchette, K. J., G. L. Allee, M. D. Newcombe, K. M. Halpin, and R. D. Boyd. 1996. Lysine requirement of the lactating primiparous sow. J. Anim. Sci. 74 (Suppl. 1):63 (Abstr.). Wang, T. C., and M. F. Fuller. 1989. The optimum dietary amino acid pattern for growing pigs. 1. Experiments by amino acid deletion. Br. J. Nutr. 62:77–89.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

4

Minerals

Pigs have a dietary requirement for certain inorganic elements. These include calcium, chlorine, copper, iodine, iron, magnesium, manganese, phosphorus, potassium, selenium, sodium, sulfur, and zinc. Chromium is now recognized as an essential mineral (National Research Council, 1997), but a quantitative requirement has not been established. Cobalt also is required in the synthesis of vitamin B12. Pigs may also require other trace elements (i.e., arsenic, boron, bromine, fluorine, molybdenum, nickel, silicon, tin, and vanadium) which have been shown to have a physiological role in one or more species (Underwood, 1977; Nielsen, 1984). These elements are required at such low levels, however, that their dietary essentiality has not been proven. The functions of these inorganic elements are extremely diverse. They range from structural functions in some tissues to a wide variety of regulatory functions in other tissues. Most pigs are now raised in confinement, without access to soil or forage; this rearing environment may increase the need for mineral supplementation. Suggested minimum requirements for the individual elements at various stages of the life cycle are given in tables provided in Chapter 10. Meeting the mineral requirements will be influenced by the bioavailabilities of minerals in feed ingredients. The subject of bioavailability of minerals was included in a recent book, Bioavailability of Nutrients for Animals, edited by Ammerman, Baker, and Lewis (1995). Several minerals, including antimony, arsenic, cadmium, fluorine, lead, and mercury, can be toxic to swine (Carson, 1986). The toxicities and tolerances of essential and other mineral elements are described in detail in Mineral Tolerance of Domestic Animals (National Research Council, 1980).

form many other physiologic functions (Hays, 1976; Peo, 1976, 1991; Kornegay, 1985). Peo (1991) indicated that adequate calcium and phosphorus nutrition for all classes of swine is dependent upon: (1) an adequate supply of each element in an available form in the diet, (2) a suitable ratio of available calcium and phosphorus in the diet, and (3) the presence of adequate vitamin D. A wide calcium-tophosphorus ratio lowers phosphorus absorption, resulting in reduced growth and bone calcification, especially if the diet is marginal in phosphorus (Peo et al., 1969; Vipperman et al., 1974; Doige et al., 1975; van Kempen et al., 1976; Reinhart and Mahan, 1986; Hall et al., 1991; Wilde and Jourquin, 1992; Eeckhout et al., 1995; Qian et al., 1996). The ratio is less critical if the diet contains excess phosphorus (Prince et al., 1984; Hall et al., 1991). A suggested ratio of total calcium-to-total phosphorus for grain–soybean meal diets is between 1:1 and 1.25:1. When based on available phosphorus, the ratio is between 2:1 and 3:1 (Jongbloed, 1987; Ketaren et al., 1989; Qian et al., 1996). A narrower calcium-to-phosphorus ratio, whether total or available phosphorus, probably results in more efficient utilization of phosphorus. An adequate amount of vitamin D is also necessary for proper metabolism of calcium and phosphorus, but a very high level of vitamin D can mobilize excessive amounts of calcium and phosphorus from bones (Hancock et al., 1986; Jongbloed, 1987). A considerable amount of research has been conducted to determine the calcium and phosphorus requirements of weanling pigs (Rutledge et al., 1961; Combs and Wallace, 1962; Combs et al., 1962, 1966; Miller et al., 1962, 1964a,b, 1965b,c,d; Menehan et al., 1963; Zimmerman et al., 1963; Blair and Benzie, 1964; Mudd et al., 1969; Coalson et al., 1972, 1974; Mahan et al., 1980; Mahan, 1982) and growingfinishing swine (Chapman et al., 1962; Libal et al., 1969; Cromwell et al., 1970, 1972b; Stockland and Blaylock, 1973; Doige et al., 1975; Pond et al., 1975, 1978; Fammatre et al., 1977; Kornegay and Thomas, 1981; Thomas and Kornegay, 1981; Maxson and Mahan, 1983; Combs et al.,

MACRO MINERALS Calcium and Phosphorus Calcium and phosphorus play a major role in the development and maintenance of the skeletal system and per-

47

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

48

Nutrient Requirements of Swine

1991a,b). The estimated dietary requirements for calcium and phosphorus for maximum growth rate and feed efficiency of pigs from 3 to 120 kg are given in Chapter 10, Tables 10-5 and 10-6. The requirements for total calcium and total phosphorus are based on a fortified, corn–soybean meal diet and take into account the fact that some of the phosphorus in feedstuffs of plant origin is unavailable. Estimates of the requirements for available phosphorus for the maximum rate and efficiency of gain are also presented in Chapter 10, Tables 10-5 and 10-6. Higher dietary concentrations of calcium and phosphorus may be required if feed intake is low. The levels of calcium and phosphorus that result in maximum growth rate are not necessarily adequate for maximum bone mineralization. The requirements for maximizing bone strength and bone-ash content are at least 0.1 percentage unit higher than the requirements for maximum rate and efficiency of gain (Cromwell et al., 1970, 1972b; Mahan et al., 1980; Crenshaw et al., 1981; Kornegay and Thomas, 1981; Mahan, 1982; Maxson and Mahan, 1983; Koch et al., 1984; Combs et al., 1991a,b). However, maximization of bone strength by feeding large amounts of calcium and phosphorus to growing pigs does not necessarily improve structural soundness (Pointillart and Gueguen, 1978; Kornegay and Thomas, 1981; Calabotta et al., 1982; Kornegay et al., 1981a,b, 1983; Breman and Aherne, 1984; Lepine et al., 1985; Eeckhout et al., 1995), nor has it been shown to be necessary for good health or longevity (Arthur et al., 1983a,b; Kornegay et al., 1984). The dietary calcium and phosphorus requirements, expressed as a percentage of the diet, may be slightly higher for gilts than for barrows (Thomas and Kornegay, 1981; Calabotta et al., 1982). Feeding of dietary levels of calcium and phosphorus sufficient to maximize bone mineralization in gilts during early growth and development was shown to improve reproductive longevity in one study (Nimmo et al., 1981a,b) but not in other studies (Arthur et al., 1983a,b; Kornegay et al., 1984). The calcium and phosphorus requirements of the developing boar are greater than those of the barrow and gilt (Cromwell et al., 1979; Hickman et al., 1983; Kesel et al., 1983; Hansen et al., 1987). Pigs possessing a high lean growth rate do not seem to have a higher dietary requirement for calcium and phosphorus as compared with pigs having a moderate lean growth rate, according to a study by Bertram et al. (1994). However, when the lean growth rate is increased by treating pigs with porcine somatotropin, the dietary requirement, expressed as percentage of the diet, increases due to the reduced daily feed intake resulting from porcine somatotropin treatment (Weeden et al., 1993a,b; Carter and Cromwell, 1998a,b). There is also strong evidence that porcine somatotropin–treated pigs require greater daily amounts of calcium and phosphorus to maximize growth

performance, bone mineralization, and carcass leanness than untreated pigs (Carter and Cromwell, 1998a,b). Kornegay et al. (1973), Harmon et al. (1974b, 1975), Monegue et al. (1980), Nimmo et al. (1981a,b), Mahan and Fetter (1982), Arthur et al. (1983a,b), Grandhi and Strain (1983), Kornegay and Kite (1983), and Maxson and Mahan (1986) investigated the calcium and phosphorus requirements of breeding swine. During pregnancy, the physiological requirements for calcium and phosphorus increase in proportion to the need for fetal growth and reach a maximum in late gestation. During lactation, the requirements are affected by the level of milk production by the sow. Generally, the requirements for calcium and phosphorus are based on a feeding level of 1.8 to 2.0 kg of feed/day during gestation and 5 to 6 kg of feed/day during lactation. If sows are fed less than 1.8 kg of feed during gestation, the diet should be formulated to contain sufficient concentrations of calcium and phosphorus to meet the daily requirements. The voluntary feed intake of lactating sows may be reduced by high environmental temperatures. In this circumstance, assuming that milk production is not decreased, the lactation diet should be formulated to meet the daily needs of calcium and phosphorus. Adequate calcium and phosphorus intakes are more critical in first-parity sows than in mature sows (Giesemann et al., 1992a,b). The form in which phosphorus exists in natural feedstuffs influences the efficiency of its utilization. In cereal grains, grain by-products, and oilseed meals, about 60 to 75 percent of the phosphorus is organically bound in the form of phytate (Nelson et al., 1968; Lolas et al., 1976), which is poorly available to the pig (Taylor, 1965; Peeler, 1972; Cromwell, 1979). The biological availability of phosphorus in cereal grains is variable (Cromwell et al., 1972a, 1974), ranging from less than 15 percent in corn (Bayley and Thomson, 1969; Miracle et al., 1977; Calvert et al., 1978; Hayes et al., 1979; Stober et al., 1979; Trotter and Allee, 1979a,b; Huang and Allee, 1981; Ross et al., 1983) to approximately 50 percent in wheat (Miracle et al., 1977; Hayes et al., 1979; Trotter and Allee, 1979a; Cromwell et al., 1985; Cromwell, 1992). The greater availability of phosphorus in wheat and wheat by-products (Stober et al., 1980b; Hew et al., 1982) is attributed to the presence of a naturally occurring phytase enzyme in wheat (McCance and Widdowson, 1944; Mollgaard, 1946; Pointillart et al., 1984). The phosphorus in high-moisture corn or grain sorghum is considerably more available than that in dry grain (Trotter and Allee, 1979b; Boyd et al., 1983; Ross et al., 1983). The phosphorus in low-phytic acid corn (modified by the mutant lpa1 gene) is relatively high (77 percent) in its bioavailability (Cromwell et al., 1998). The phosphorus in oilseed meals also has a low bioavailability (Tonroy et al., 1973; Miracle et al., 1977; Trotter and Allee, 1979a; Stober et al., 1980a; Harrold, 1981; Ross

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals et al., 1982; Cromwell, 1992). In contrast, the phosphorus in protein sources of animal origin is largely inorganic, and most animal protein sources (including milk and blood byproducts) have a high phosphorus bioavailability (Cromwell et al., 1976; Hew et al., 1982; Coffey and Cromwell, 1993). The bioavailability of phosphorus in meat and bone meal is variable. Earlier studies indicated that the bioavailability of phosphorus in meat and bone meal was somewhat lower (67%) than in most other animal sources (Cromwell, 1992), but more recent studies showed a relatively high bioavailability (90%) (Traylor and Cromwell, 1998). The phosphorus in dehydrated alfalfa meal is highly available (Cromwell et al., 1983). Steam pelleting has been shown to improve the bioavailability of phytate phosphorus in some studies (Bayley and Thompson, 1969; Bayley et al., 1975) but not in others (Trotter and Allee, 1979c; Corley et al., 1980; Ross et al., 1983). Estimates of relative phosphorus bioavailability in common feed ingredients for pigs are given in Chapter 11, Table 11-1. Microbial phytase supplementation of high-phytate, cereal grain– oilseed meal diets can result in major improvements in bioavailability of phytate phosphorus (Nasi, 1990; Simmons et al., 1990; Jongbloed et al., 1992; Pallauf et al., 1992a,b; Cromwell et al., 1993b, 1995; Lei et al., 1993a). As a result, the dietary level of phosphorus can be reduced, thereby lowering phosphorus excretion by 30 to 60 percent (see Chapter 8). The magnitude of the response to microbial phytase is influenced by the dietary level of available and total phosphorus (including phytate phosphorus), the amount of supplemental phytase, the calcium-to-phosphorus ratio (or level of calcium), and the level of vitamin D (Jongbloed et al., 1993; Lei et al., 1994; Kornegay, 1996). Microbial phytase also improves the bioavailability of calcium (Lei et al., 1993a; Mroz et al., 1994; Pallauf et al., 1992b; Young et al., 1993; Radcliffe et al., 1995) and zinc (Pallauf et al., 1992a, 1994a,b; Lei et al., 1993b) and has been reported to improve the digestibility of dietary protein (Ketaren et al., 1993; Kornegay and Qian, 1994; Mroz et al., 1994; Kemme et al., 1995; Biehl and Baker, 1996). Pelleting of diets can reduce or destroy phytase activity because of the temperature increases that occur during the pelleting process. Loss of phytase activity has been reported when temperatures exceed 60°C (Jongbloed and Kemme, 1990; Nunes, 1993); such a loss can result in reduced digestibility of phosphorus and calcium (Jongbloed and Kemme, 1990). The phosphorus in inorganic phosphorus supplements also varies in bioavailability. The phosphorus in ammonium, calcium, and sodium phosphates is highly available (Kornegay, 1972b; Hays, 1976; Clawson and Armstrong, 1981; Partridge, 1981; Tunmire et al., 1983; Cromwell et al., 1987; Cromwell, 1992). The phosphorus in steamed bone meal is less available than that in mono-dicalcium phosphate (Cromwell, 1992). The phosphorus in defluori-

49

nated rock phosphate is generally less available than in monocalcium phosphate or monosodium phosphate (Cromwell, 1992; Coffey et al., 1994) but can vary depending on source and processing (Kornegay and Radcliffe, 1997). The phosphorus in high-fluorine rock phosphates, ¸ soft phosphate, colloidal clay, and Curacao phosphate is poorly available (Chapman et al., 1955; Plumlee et al., 1958; Harmon et al., 1974b; Hays, 1976; Peo et al., 1982a,b). Estimates of the bioavailability of phosphorus in phosphorus supplements are given in Chapter 11 (Table 11-8). Little is known about the availability of calcium in natural feedstuffs. Because of the phytic acid content, the bioavailability of calcium in cereal grain–based diets, alfalfa, and various grasses and hays is relatively low (Soares, 1995). However, most feedstuffs contribute so little calcium to the diet that bioavailability of the calcium is of little consequence. The calcium in calcitic limestone, gypsum, oystershell flour, aragonite, and marble dust is highly available (Pond et al., 1981; Ross et al., 1984), but the calcium in dolomitic limestone is only 50 to 75 percent available (Ross et al., 1984). Particle size (up to 0.5 mm in diameter) appears to have little effect on calcium availability (Ross et al., 1984). Pig data are not available, but on the basis of poultry data, the calcium in dicalcium phosphate, tricalcium phosphate, defluorinated phosphate, calcium gluconate, calcium sulfate, and bone meal is highly available, generally 90 to 100 percent, when compared with the calcium in calcium carbonate (Baker, 1991; Soares, 1995). Signs of calcium or phosphorus deficiency are similar to those of vitamin D deficiency. They include depressed growth and poor bone mineralization, resulting in rickets in young pigs and osteomalacia in older swine. A common problem of calcium- or phosphorus-deficient sows is a paralysis of the hind legs, called posterior paralysis. The problem occurs most frequently in sows producing high levels of milk toward the end or just after the termination of lactation. Excess levels of calcium and phosphorus may reduce performance of pigs (Hall et al., 1991; Reinhart and Mahan, 1986), and the effect is greater when the calcium:phosphorus ratio is increased. Excess calcium not only decreases the utilization of phosphorus but also increases the pig’s requirement for zinc in the presence of phytate (Luecke et al., 1956; Whiting and Bezeau, 1958; Morgan et al., 1969; Oberleas, 1983). When the molar ratio of cations (zinc and calcium) was 2:1 or 3:1 with phytate, the formation of an insoluble complex was much greater (Oberleas and Harland, 1996). Excess calcium also increases the requirement for vitamin K (Hall et al., 1991). Sodium and Chlorine Sodium and chlorine (chloride) are the principal extracellular cation and anion, respectively, in the body. Chloride is the chief anion in gastric juice.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

50

Nutrient Requirements of Swine

The dietary sodium requirement of growing-finishing pigs is no greater than 0.08 to 0.10 percent of the diet (Meyer et al., 1950; Alcantara et al., 1980; Cromwell et al., 1981a; Froseth et al., 1982a; Honeyfield and Froseth, 1985; Honeyfield et al., 1985; Kornegay et al., 1991). The dietary chlorine (chloride) requirement is less well defined but is probably no higher than 0.08 percent for the growing pig (Froseth et al., 1982a; Honeyfield and Froseth, 1985; Honeyfield et al., 1985). A level of 0.20 to 0.25 percent added sodium chloride will meet the dietary sodium and chlorine requirements of growing-finishing pigs fed a corn– soybean meal diet (Hagsten and Perry, 1976a,b; Hagsten et al., 1976). Mahan et al. (1996a,b) recently reported that weanling pigs fed diets containing dried whey or dried plasma (both are relatively high in sodium) responded to added sodium as sodium chloride or sodium phosphate and to added chloride as hydrochloric acid. Their results indicate that early-weaned pigs require more sodium and chlorine than previously thought. Thus, the estimated dietary sodium and chloride requirements have been increased to 0.25 percent of each from 3 to 5 kg, to 0.20 percent of each from 5 to 10 kg, and to 0.15 percent of each from 10 to 20 kg body weight. The sodium and chlorine requirements of breeding animals are not well established. The results of one study suggested that 0.3 percent dietary sodium chloride (0.12 percent sodium) was not sufficient for pregnant sows (Friend and Wolynetz, 1981). In a regional study, pig birth weights and weaning weights were reduced when sodium chloride was reduced from 0.50 to 0.25 percent during gestation and lactation for two or more parities (Cromwell et al., 1989a). Based upon the sodium content of sow’s milk, which is 0.03 to 0.04 percent (Agricultural Research Council, 1981), the dietary sodium requirement should be about 0.05 percentage unit greater during lactation than during gestation. Until more definitive information is available, sodium chloride additions of 0.4 percent to gestation diets and 0.5 percent to lactation diets are suggested. The availability of sodium and chloride in most feed ingredients is believed to be 90 to 100 percent (Miller, 1980). The sodium in water, which in coastal regions can be as high as 184 mg/L, and in defluorinated phosphate is highly available for pigs (Kornegay et al., 1991). A deficiency of sodium or chloride reduces the rate and efficiency of growth in pigs. In contrast, swine can tolerate high dietary levels of sodium chloride (National Research Council, 1980), provided they have access to ample nonsaline drinking water. If non-saline water is limited or if the level of sodium chloride in water is high, toxicity can result. The high sodium ion concentration is responsible for adverse physiological reactions, apparently because of a disturbance in water balance. The signs of sodium toxicity include nervousness, weakness, staggering, epileptic sei-

zures, paralysis, and death (Bohstedt and Grummer, 1954; Carson, 1986). Sodium, potassium, and chloride are the primary dietary ions that influence the electrolyte balance and acid-base status of animals. Under most circumstances, dietary mineral balance is expressed as milliequivalents (mEq) of sodium plus potassium minus chloride ions (Na ` K 1 Cl) (Mongin, 1981) and is often referred to as electrolyte balance. Patience and Wolynetz (1990) suggested that calcium, magnesium, sulfur, and phosphorus ions should also be included in the calculation of electrolyte balance. The optimal electrolyte balance in the diet for pigs is 250 mEq of excess cations (Na ` K 1 Cl)/kg of diet according to Austic and Calvert (1981), Golz and Crenshaw (1990), and Haydon et al. (1993); however, optimal growth has been found to occur over the range of 0 to 600 mEq/kg of diet (Patience et al., 1987; Kornegay et al., 1994). If a deficiency of sodium, potassium, or chloride occurs in the diet, then the relationship, Na ` K 1 Cl, does not accurately predict dietary levels for optimum growth (Mongin, 1981).

Magnesium Magnesium is a cofactor in many enzyme systems and is a constituent of bone. The magnesium requirement of artificially reared pigs fed milk-based semipurified diets is between 300 and 500 mg/kg of diet (Mayo et al., 1959; Bartley et al., 1961; Miller et al., 1965a,c,d). Milk contains adequate magnesium to meet the requirement of suckling pigs (Miller et al., 1965c,d). The magnesium requirement of weanling-growing-finishing swine is probably not higher than that of the young pig. The magnesium in a corn– soybean meal diet (0.14 to 0.18 percent) is apparently adequate (Svajgr et al., 1969; Krider et al., 1975), although some research suggests that the magnesium in natural ingredients is only 50 to 60 percent available to the pig (Miller, 1980; Nuoranne et al., 1980). The magnesium requirement of breeding animals is not well established. Harmon et al. (1976) fed semipurified diets containing 0.04 and 0.09 percent magnesium to sows during gestation, followed by 0.015 and 0.065 percent magnesium during lactation. They observed no difference in reproductive or lactational performance. However, in a balance study, sows fed the low level of magnesium during lactation were in negative magnesium balance. In order of appearance, signs of magnesium deficiency include hyperirritability, muscular twitching, reluctance to stand, weak pasterns, loss of equilibrium, and tetany followed by death (Mayo et al., 1959; Miller et al., 1965c). The toxic level of magnesium is not known. The maximum tolerable level for swine is approximately 0.3 percent (National Research Council, 1980).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals

51

Potassium

Sulfur

Potassium is the third most abundant mineral in the body of the pig, surpassed only by calcium and phosphorus (Manners and McCrea, 1964), and is the most abundant mineral in muscle tissue (Stant et al., 1969). Potassium is involved in electrolyte balance and neuromuscular function. It also serves as the monovalent cation to balance anions intracellularly, as part of the sodium potassium pump physiological mechanism. The dietary potassium requirement of pigs from 1 to 4 kg body weight is estimated to be between 0.27 and 0.39 percent (Manners and McCrea, 1964); from 5 to 10 kg, 0.26 to 0.33 percent (Jensen et al., 1961; Combs et al., 1985); at 16 kg, 0.23 to 0.28 percent (Meyer et al., 1950); and from 20 to 35 kg, less than 0.15 percent (Hughes and Ittner, 1942; Mraz et al., 1958). No estimates are available for finishing or breeding pigs. The content of potassium in most practical diets is normally adequate to meet these requirements for all classes of swine. The potassium in corn and soybean meal is 90 to 97 percent available (Combs and Miller, 1985). Dietary potassium is interrelated with dietary sodium and chloride. Increasing dietary chloride from 0.03 to 0.60 percent in purified diets depressed growth rate of young pigs when the diet contained 0.1 percent potassium, but it increased growth rate when the diet contained 1.1 percent potassium (Golz and Crenshaw, 1990). The interactive effect of dietary potassium and chloride seems to be an indirect effect on the excretion and retention of additional cations and anions, particularly ammonium and phosphate. The effects on growth are mediated via mechanisms involving renal ammonium ion metabolism (Golz and Crenshaw, 1991). Potassium has been shown to spare lysine in the chick, but a similar response has not been demonstrated consistently in the pig (Leibholz et al., 1966; Madubuike et al., 1980; Austic and Calvert, 1981; Miller et al., 1981b; Wahlstrom and Libal, 1981; Froseth et al., 1982b,c; Miller and Froseth, 1982; Zimmerman, 1982; Mijada and Cline, 1983). Madubuike and Austic (1989) suggested that this inconsistency may be related to the lysine adequacy of the pig diet. Signs of potassium deficiency include anorexia, rough hair coat, emaciation, inactivity, and ataxia (Jensen et al., 1961). Electrocardiograms of potassium-deficient pigs showed reduced heart rate and increased electrocardial intervals (Cox et al., 1966). Necropsy of affected pigs revealed no unique gross pathology. The toxic level of potassium is not well established. Pigs can tolerate up to 10 times the potassium requirement if plenty of drinking water is provided (Farries, 1958). Intravenous infusion of potassium chloride in pigs resulted in abnormal electrocardiograms (Coulter and Swenson, 1970).

Sulfur is an essential element. The sulfur provided by the sulfur-containing amino acids seems adequate to meet the pig’s needs for synthesis of sulfur-containing compounds, such as taurine, glutathione, lapoic acid, and chondroitin sulfate. Additions of inorganic sulfate to low-protein diets have not been beneficial (Miller, 1975; Baker, 1977).

MICRO/TRACE MINERALS Chromium Chromium is involved in carbohydrate, lipid, protein, and nucleic acid metabolism (Nielsen, 1994). Although the specific function of chromium is unknown, it is believed to work as a cofactor with insulin (White et al., 1993). A glucose tolerance factor containing chromium potentiated insulin activity and was biologically active (Steele et al., 1977). Increased insulin sensitivity was reported for pigs fed chromium picolinate (Amoikon et al., 1995). Chromium added as chromium picolinate was reported by Evock-Clover et al. (1993) to lower serum insulin and glucose concentrations in growing pigs (30 to 60 kg). However, in other studies serum glucose concentrations were not influenced by feeding chromium (Page et al., 1993; Amoikon et al., 1995; Lindemann et al., 1995a). Lindemann et al. (1995a) reported that postfeeding serum insulin values and insulin-to-glucose ratios were lower for fasted gestating sows fed chromium picolinate than for fasted control sows. Amoikon et al. (1995) also reported that when pigs were fed chromium picolinate, the fasting plasma insulin value was reduced; the clearance of glucose after an intravenous glucose tolerance test and insulin challenge test was increased; and the half-life of glucose was decreased. Chromium, especially inorganic forms, is poorly absorbed from the gastrointestinal tract. The amount of inorganic chromium absorbed ranges from 0.4 to 3 percent according to a review by Anderson (1987). Some organic forms are better absorbed than inorganic forms. The absorption by humans of chromium from chromium picolinate was low; 0.7 to 1.7 percent in one study (Clancy et al., 1994) and 1.5 to 5.2 percent in another (Gargas et al., 1994). Ward et al. (1995) evaluated several forms of chromium (chloride, acetate, oxalate, nicotinate, two sources of picolinate, and nicotinate-glycine-cysteine-glutamate) that were fed to supply 200 ppb chromium, but found that serum metabolites and hormone values were not affected by any of the forms of chromium. Also, chromium chloride (5 or 25 ppb chromium) or chromium picolinate (200 or 400 ppb chromium) did not affect serum metabolites in a study by Mooney and Cromwell (1997).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

52

Nutrient Requirements of Swine

Recent interest has focused on the potential use of the organic chromium complex, chromium picolinate, to increase carcass leanness. Positive responses were reported by Page et al. (1993), Lindemann et al. (1995b), Harper et al. (1995), Boleman et al. (1995), and Mooney and Cromwell (1995; 1997). However, others reported no responses in carcass leanness to supplemental chromium in this form (Ward et al., 1995; Harris et al., 1995; Mooney and Cromwell, 1996). The lack of a consistent response may be related to chromium levels of diets, form of chromium, chromium status of pig, and amino acid levels of diet (White et al., 1993; Lindemann et al., 1995b). The total chromium content of a corn–soybean diet can range from 750 to 1,500 ppb, but most of this is probably unavailable. Larger litters at birth have been reported for sows fed supplemental chromium picolinate (Lindemann et al., 1995a,b). In one large trial, farrowing rate was increased when first and second parity sows were fed 200 ppb chromium as chromium picolinate beginning on the day after breeding through farrowing, but total and live pigs born were not affected by treatments (Campbell, 1996). In a second trial, multiparous sows were fed 200 ppb chromium as chromium picolinate for the first 35 days after breeding; in a third trial they were fed the same amount for 28 days prior to farrowing or for 28 days prior to farrowing through lactation and for 35 days after breeding. The supplementation had little effect on any measure of fertility or fecundity. Additional research is required to elucidate the role of chromium in swine. The inconsistency of biological responses to chromium could be related to the bioavailability of the chromium found in traditional feed ingredients, the duration of feeding, and the chromium status of the pigs. Trivalent and hexavalent are the two most common forms of chromium; both are stable. Hexavalent chromium is much more toxic than trivalent chromium, which is believed to be the essential trace mineral (Anderson, 1987; Mertz, 1993). Maximum tolerable dietary levels for domestic animals were set at 3,000 ppm chromium as the oxide and 1,000 ppm as the chloride (National Research Council, 1980). A detailed discussion of tolerance concentration for chromium in animals can be found in Mineral Tolerance of Domestic Animals (National Research Council, 1980). The results of a recent in vitro study with Chinese hamster ovary cells indicate some chromosome damage after treatment with soluble doses of 0.05, 0.10, 0.50, and 1.0 mM of chromium picolinate (Stearns et al., 1996). Chromium nicotinate, nicotinic acid, and trivalent chromium chloride hexahydrate did not produce chromosome damage at equivalent nontoxic doses. These results suggest the need for further investigations of the long-term effects of supplemental chromium. No quantitative estimate of the chromium requirement has been estimated for pigs. A recent

review on chromium was published by the National Research Council (1997).

Cobalt Cobalt is a component of vitamin B12 (Rickes et al., 1948). There is no evidence that pigs have an absolute requirement for cobalt, other than for its role in vitamin B12. Cobalt can substitute for zinc in the enzyme carboxypeptidase and for part of the zinc in the enzyme alkaline phosphatase. Hoekstra (1970) and Chung et al. (1976) have shown that supplemental cobalt prevents lesions associated with a zinc deficiency. Dietary cobalt can only be used by the intestinal microflora of the pig to synthesize some vitamin B12. Intestinal synthesis assumes greater importance if dietary vitamin B12 is limiting (Klosterman et al., 1950; Robinson, 1950; Kline et al., 1954). The use of supplemental vitamin B12 in practical diets is a routine practice. A level of 400 ppm cobalt was toxic to the young pig and may cause anorexia, stiff-leggedness, humped back, incoordination, muscle tremors, and anemia (Huck and Clawson, 1976). Cobalt concentration in the kidney and liver increased linearly and growth decreased linearly over a 4- to 5-week period as 0, 150, and 300 ppm cobalt were added to a basal diet containing ,2 ppm cobalt (Kornegay et al., 1995). The maximum tolerance level for weanling pigs is ,150 ppm of diet. Selenium, vitamin E, and cysteine provide some protection against toxicity from excessive levels of dietary cobalt (Van Vleet et al., 1977; Southern and Baker, 1981), but growth-stimulating levels of copper may aggravate the growth depression caused by cobalt (Kornegay et al., 1995).

Copper The pig requires copper for the synthesis of hemoglobin and for the synthesis and activation of several oxidative enzymes necessary for normal metabolism (Miller et al., 1979). A level of 5 to 6 ppm in the diet is adequate for the neonatal pig (Okonkwo et al., 1979; Hill et al., 1983a). The requirement for later stages of growth is probably no greater than 5 to 6 ppm. Definitive information on requirements during gestation and lactation are scarce. Lillie and Frobish (1978) suggested that 60 ppm of copper fed to sows improved pig weights at birth and at weaning, but this response may have resulted from the pharmacological effect of high dietary copper. Kirchgessner et al. (1980) found that pregnant sows fed 2 ppm of copper had reduced ceruloplasmin and farrowed more stillborn pigs than sows fed 9.5 ppm of copper. In a balance study, Kirchgessner et al. (1981) estimated the copper requirement of pregnant sows at 6 ppm.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Copper salts with high biological availabilities include the sulfate, carbonate, and chloride salts (Miller, 1980; Cromwell et al., 1998). The copper in cupric sulfide and cupric oxide is poorly available to the pig (Cromwell et al., 1978, 1989b; Sazzad et al., 1993). Organic complexes of copper appear to have equal bioavailability to copper sulfate in several trials (Bunch et al., 1965; Zoubek et al., 1975; Stansbury et al., 1990; Coffey et al., 1994; Apgar et al., 1995). However, in two trials reported by Coffey et al. (1994) and Zhou et al. (1994a), growth performance was greater in pigs fed growth-promotion levels of copper from a copper lysine complex than those fed copper sulfate. A deficiency of copper leads to poor iron mobilization; abnormal hemopoiesis; and poor keratinization and synthesis of collagen, elastin, and myelin. Copper deficiency signs include a microcytic, hypochromic anemia; bowing of the legs; spontaneous fractures; cardiac and vascular disorders; and depigmentation (Hart et al., 1929; Elvehjem and Hart, 1932; Teague and Carpenter, 1951; Follis et al., 1955; Carter et al., 1959; Carnes et al., 1961; Hill et al., 1983a). Copper may be toxic when dietary levels in excess of 250 ppm are fed for extended periods of time (National Research Council, 1980). Toxicity signs include depressed hemoglobin levels and jaundice, which are the results of excessive copper accumulation in the liver and other vital organs. Reduced dietary levels of zinc and iron or high levels of dietary calcium accentuate copper toxicity (Suttle and Mills, 1966a,b; Hedges and Kornegay, 1973; Prince et al., 1984). When fed at 100 to 250 ppm, copper (as copper sulfate) stimulates growth in pigs (Barber et al., 1955b; Braude, 1967, 1975; Wallace, 1967; Cromwell et al., 1981b; Kornegay et al., 1989; Cromwell, 1997). The growth response to copper in young pigs is independent of, and in addition to, the growth response to other antibacterial agents (Stahly et al., 1980; Roof and Mahan, 1982; Edmonds et al., 1985; Cromwell 1997). The response to high levels of copper may be enhanced by added fat (Dove and Haydon, 1992; Dove, 1993a, 1995). The continuous feeding of high copper levels to sows for up to six consecutive gestation–lactation cycles did not have any apparent negative effects on reproductive performance, in spite of rather large increases in liver and kidney copper concentrations (Cromwell et al., 1993a). In fact, birth and weaning weights were greater in pigs from sows fed high copper. Improved weight gain of suckling pigs was also observed by Lillie and Frobish (1978), but other studies in which copper was fed during late gestation and lactation (Thacker, 1991) or during lactation (Roos and Easter, 1986; Dove, 1993b) showed no response to added copper in weight gain of suckling pigs. The mechanisms through which beneficial effects from copper are observed are unknown. The growth-stimulating action of dietary copper has been attributed to its antimicrobial actions (Fuller et al., 1960); however, evidence

53

supporting this hypothesis is lacking. A correlation between the availability of copper and the growth-promoting action of copper has been observed (Bowland et al., 1961; Cromwell et al., 1989b). Zhou et al. (1994b) reported that both body weight gain and serum mitogenic activity were stimulated in young pigs given intravenous injections of copper histidinate every other day for 18 days. Because the gastrointestinal tract was bypassed in this study, these results suggest that copper can act systemically to promote growth. Feeding 250 ppm copper stimulated lipase and phospholipase A activities and led to an improvement of dietary fat digestibility in weaning pigs (Luo and Dove, 1996). High dietary levels of copper increase fecal copper excretion, but the form of copper in pig feces is poorly bioavailable to chickens and sheep (Prince et al., 1975; Izquierdo and Baker, 1986).

Iodine The majority of the iodine in swine is present in the thyroid gland, where it exists as a component of mono-, di-, tri-, and tetraiodothyronine (thyroxine). These hormones are important in the regulation of metabolic rate. Hart and Steenbock (1918), Kalkus (1920), and Welch (1928) demonstrated that hypothyroidism existed in swine raised in the northwestern United States and the Great Lakes region because of iodine-deficient feedstuffs produced on low-iodine soil. The dietary iodine requirement is not well established. The requirement is increased by goitrogens, which are present in certain feedstuffs, including rapeseed, linseed, lentils, peanuts, and soybeans (McCarrison, 1933; Underwood, 1977). A level of 0.14 ppm of iodine in a corn– soybean meal diet is adequate to prevent thyroid hypertrophy in growing pigs (Cromwell et al., 1975). A level of 0.35 ppm of added iodine prevented iodine deficiency in sows (Andrews et al., 1948). Calcium iodate, potassium iodate, and pentacalcium orthoperiodate are nutritionally available forms of iodine and are more stable in salt mixtures than are sodium iodide or potassium iodide (Kuhajek and Andelfinger, 1970). The incorporation of iodized salt (0.007 percent iodine), at a level of 0.2 percent of the diet, provides sufficient iodine (0.14 ppm) to meet the needs of growing pigs fed grain–soybean meal diets. A severe iodine deficiency causes pigs to be stunted and lethargic and to have an enlarged thyroid (Beeson et al., 1947; Braude and Cotchin, 1949; Sihombing et al., 1974). Sows fed iodine-deficient, goitrogenic diets farrow weak or dead pigs that are hairless, show symptoms of myxedema, and have an enlarged, hemorrhagic thyroid (Hart and Steenbock, 1918; Slatter, 1955; Devilat and Skoknic, 1971).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

54

Nutrient Requirements of Swine

A dietary iodine level of 800 ppm depressed growth, hemoglobin level, and liver iron concentration in growing pigs (Newton and Clawson, 1974). During lactation and the last 30 days of gestation, as much as 1,500 to 2,500 ppm of iodine was not harmful to sows (Arrington et al., 1965). Iron Iron is required as a component of hemoglobin in red blood cells. Iron also is found in muscle as myoglobin, in serum as transferrin, in the placenta as uteroferrin, in milk as lactoferrin, and in the liver as ferritin and hemosiderin (Zimmerman, 1980; Ducsay et al., 1984). It also plays an important role in the body as a constituent of several metabolic enzymes. Pigs are born with about 50 mg of iron, most of which is present as hemoglobin (Venn et al., 1947). A high level of iron fed to sows during late gestation (Brady et al., 1978) or parenteral administration of iron dextran to sows in gestation (Rydberg et al., 1959; Pond et al., 1961; Ducsay et al., 1984) does not substantially increase placental transfer of iron to fetuses. The suckling pig must retain 7 to 16 mg of iron daily, or 21 mg of iron/kg of body weight gain to maintain adequate levels of hemoglobin and storage iron (Venn et al., 1947; Braude et al., 1962). Sow’s milk contains an average of only 1 mg of iron per liter (Brady et al., 1978). Thus, pigs receiving only milk rapidly develop anemia (Hart et al., 1929; Venn et al., 1947). Feeding of high levels of various iron compounds, including iron sulfate and iron chelates, to gestating and lactating sows does not increase the iron content of milk to an extent that iron deficiency can be prevented. These levels can, however, prevent iron deficiency in suckling pigs that have access to the sow’s feces (Chaney and Barnhart, 1963; Veum et al., 1965; Spruill et al., 1971; Brady et al., 1978; Sansom and Gleed, 1981; Gleed and Sansom, 1982). The iron requirement of young pigs fed milk or purified liquid diets is 50 to 150 mg/kg of milk solids (Matrone et al., 1960; Ullrey et al., 1960; Manners and McCrea, 1964; Harmon et al., 1967; Hitchcock et al., 1974). Miller et al. (1982) suggested a requirement of 100 mg of iron/kg of milk solids for pigs raised in a conventional or germ-free environment. The iron requirement of pigs fed a dry, casein-based diet is about 50 percent higher per unit of dry matter than for those fed a similar diet in liquid form (Hitchcock et al., 1974). Numerous studies have shown the effectiveness of a single intramuscular injection of 100 to 200 mg of iron, in the form of iron dextran, iron dextrin, or gleptoferron given in the first 3 days of life (Barber et al., l955a; McDonald et al., 1955; Maner et al., 1959; Rydberg et al., 1959; Ullrey et al., 1959; Zimmerman et al., 1959; Linkenheimer et al., 1960; Wahlstrom and Juhl, 1960; Kernkamp et al., 1962; Parsons, 1979; Pollmann et al., 1983). The intestinal

mucosa of the newborn pig actively absorbs iron (Furugouri and Kawabata, 1975, 1976, 1979). Oral administration of iron from bioavailable inorganic or organic sources within the first few hours of life also will meet the iron needs of the suckling pig. However, early administration, before gut closure to large molecules, is crucial (Harmon et al., 1974a; Thoren-Tolling, 1975). An excessive level (more than 200 mg) of injectable or oral iron should be avoided because unbound serum iron encourages bacterial growth and results in increased susceptibility to infection and diarrhea (Weinberg, 1978; Klasing et al., 1980; Knight et al., 1983; Kadis et al., 1984). The postweaning dietary iron requirement is about 80 ppm (Pickett et al., 1960). In later growth and maturity, this requirement diminishes as the rate of increase in blood volume slows. Natural feed ingredients usually supply enough iron to meet postweaning requirements. Feedgrade defluorinated phosphate and dicalcium phosphate, which contain from 0.6 to 1.0 percent iron, also supply substantial amounts of iron. The iron in defluorinated phosphate is about 65 percent as available to the pig as the iron in ferrous sulfate (Kornegay, 1972a). Availability of iron from different sources varies greatly (Zimmerman, 1980). Ferrous sulfate, ferric chloride, ferric citrate, ferric choline citrate, and ferric ammonium citrate are effective in preventing iron deficiency anemia (Harmon et al., 1967; Ammerman and Miller, 1972; Ullrey et al., 1973; Miller et al., 1981a). Iron compounds with low solubility, such as ferric oxide, are ineffective (Ammerman and Miller, 1972). The bioavailability of iron in ferrous carbonate is lower and more variable than that of iron in ferrous sulfate (Harmon et al., 1969; Ammerman et al., 1974). Iron from iron methionine was 68 to 81 percent as bioavailable as that in iron sulfate (Lewis et al., 1995). Soybean meal contains 175 to 200 ppm of iron, and the bioavailability of iron in soybean meal has been estimated to be 38 percent, based on hemoglobin depletion–repletion assays in chicks (Biehl et al., 1997). The hemoglobin concentration of blood is a reliable indicator of the pig’s iron status, and it is easy to determine. Hemoglobin levels of 10 g/dL of whole blood are considered adequate. A hemoglobin level of 8 g/dL suggests borderline anemia, and a level of 7 g/dL or less represents anemia (Zimmerman, 1980). The type of anemia resulting from iron deficiency is hypochromic-microcytic anemia. Anemic pigs show evidence of poor growth, listlessness, rough hair coats, wrinkled skin, and paleness of mucous membranes. Fast-growing anemic pigs may die suddenly of anoxia. A characteristic sign is labored breathing after minimal activity or a spasmodic jerking of the diaphragm muscles, from which the term ‘‘thumps’’ arises. Necropsy findings include an enlarged and fatty liver; thin, watery blood; marked dilation of the heart; and an enlarged firm

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals spleen. Anemic pigs are more susceptible to infectious diseases (Osborne and Davis, 1968). In 3- to 10-day-old pigs, the toxic oral dose of iron from ferrous sulfate is approximately 600 mg/kg of body weight (Campbell, 1961). Clinical signs of toxicity are observed within 1 to 3 hours after iron is fed (Nilsson, 1960; Arpi and Tollerz, 1965). Lannek et al. (1962) and Patterson et al. (1967, 1969) have found that injectable iron (100 mg as iron dextran) is toxic to pigs from vitamin E–deficient dams. A dietary level of 5,000 ppm of iron produces rachitic lesions, which may be prevented by increasing the level of dietary phosphorus (O’Donovan et al., 1963; Furugouri, 1972).

Manganese Manganese functions as a component of several enzymes involved in carbohydrate, lipid, and protein metabolism. Manganese is essential for the synthesis of chondroitin sulfate, a component of mucopolysaccharides in the organic matrix of bone (Leach and Muenster, 1962). The dietary requirements for manganese are quite low (Johnson, 1944) and not well established. The manganese status of the sow affects the manganese status of the neonates, because manganese readily crosses the placenta (Newland and Davis, 1961; Gamble et al., 1971). Leibholz et al. (1962) reported that as little as 0.4 ppm of manganese is sufficient for young pigs. With manganese-depleted dams, however, the requirement for the neonates is 3 to 6 ppm (Kayongo-Male et al., 1975). A corn–soybean meal diet should contain ample manganese for normal growth and bone formation in growing-finishing pigs (Svajgr et al., 1969). Long-term feeding of a diet containing only 0.5 ppm of manganese results in abnormal skeletal growth, increased fat deposition, irregular or absent estrous cycles, resorbed fetuses, small, weak pigs at birth, and reduced milk production (Plumlee et al., 1956). On the basis of manganese retention, Kirchgessner et al. (1981) estimated the manganese requirement of pregnant sows at 25 ppm. Total litter weight at birth was less for sows fed a low-manganese, basal corn–soybean meal diet (10 ppm manganese) than for sows fed the basal diet plus 84 ppm manganese (Rheaume and Chavaz, 1989). Colostrum and milk from sows fed supplemental manganese contained a higher concentration of manganese, but retention of manganese was only numerically higher. Christianson et al. (1989, 1990) reported that birth weight of pigs was greater when sows were fed 10 or 20 ppm manganese than when they were fed 5 ppm. Also, return to estrus was improved by feeding 20 ppm manganese. On the basis of these recent studies, the manganese requirements for gestation and lactation have been increased to 20 ppm of the diet.

55

Although the toxic level of manganese is not well defined, depressed feed intake and reduced growth rates have been observed when pigs were fed 4,000 ppm of manganese (Leibholz et al., 1962). A dietary level of 2,000 ppm of manganese resulted in reduced hemoglobin levels (Matrone et al., 1959), and 500 ppm of manganese reduced growth rate and resulted in limb stiffness in growing pigs (Grummer et al., 1950). Selenium Selenium is a component of the enzyme glutathione peroxidase (Rotruck et al., 1973), which detoxifies lipid peroxides and provides protection of cellular and subcellular membranes against peroxide damage. Thus, the mutual sparing effect of selenium and vitamin E stems from their shared antiperoxidant roles. High levels of vitamin E, however, do not completely eliminate the need for selenium (Ewan et al., 1969; Bengtsson et al., 1978a,b; Hakkarainen et al., 1978). Selenium has been shown to have a function in thyroid metabolism, because iodothyronine 58-deiodinase has been identified as a selenoprotein (Arthur, 1994). The dietary requirement for selenium ranges from 0.3 ppm for weanling pigs to 0.15 ppm for finishing pigs and sows (Groce et al., 1971, 1973a,b; Ku et al., 1973; Mahan et al., 1973; Ullrey, 1974; Young et al., 1976; Glienke and Ewan, 1977; Wilkinson et al., 1977a,b; Mahan and Moxon, 1978a,b, 1984; Piatkowski et al., 1979; Meyer et al., 1981). The requirement for selenium is influenced by dietary phosphorus level (Lowry et al., 1985b) but not dietary calcium level (Lowry et al., 1985a). Several forms of selenium, including selenium-enriched yeast, sodium selenite, and sodium selenate, are effective in meeting the dietary requirement (Mahan and Magee, 1991; Suomi and Alaviuhkola, 1992; Mahan and Parrett, 1996; Mahan and Kim, 1996). The selenium status of the dam influences reproductive performance and the selenium requirement of suckling and weanling pigs (Van Vleet et al., 1973; Mahan et al., 1977; Piatkowski et al., 1979; Chavez, 1985; Ramisz et al., 1993). Total body retention of selenium, as well as serum and tissue levels of selenium in growing, finishing, and reproducing gilts and their suckling progeny, increased as the dietary level of selenium increased (0.1 to 0.3 or 0.5 ppm); the amount of selenium retained and stored was usually greater at the various selenium levels when the effects of a selenium-enriched yeast source were compared with those produced by sodium selenite (Mahan, 1995; Mahan and Kim, 1996; Mahan and Parrett, 1996). In reproducing gilts, serum glutathione peroxidase activity was not improved beyond 0.1 ppm selenium, and the increase in activity was similar for selenium-enriched yeast and sodium selenite (Mahan and Kim, 1996). In growing-finishing pigs, serum selenium concentration and serum glutathione peroxidase activity reached a plateau at a dietary level of 0.1

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

56

Nutrient Requirements of Swine

ppm selenium for selenium-enriched yeast and sodium selenite, but the magnitude of the response was lower for the yeast than for the sodium selenite, which suggests that the selenium-enriched yeast product was less biologically available than sodium selenite (Mahan and Parrett, 1996). About 50 percent of the selenium in the selenium-enriched yeast product was suggested to be selenomethionine, with the remainder in one of several seleno-amino acids or as their analogs (Mahan, 1995). Certain soils of the United States and Canada are low in selenium. When diets consist exclusively of ingredients grown in such regions, selenium will be deficient unless supplemental selenium is added (Grant et al., 1961; Trapp et al., 1970; Ewan, 1971; Groce et al., 1971; Sharp et al., 1972a,b; Ku et al., 1973; Mahan et al., 1973, 1974; Diehl et al., 1975; Doornenbal, 1975; Piper et al., 1975; Wilkinson et al., 1977b; Bengtsson et al., 1978b). Environmental stress may increase the incidence and degree of selenium deficiency (Michel et al., 1969; Mahan et al., 1975). In 1974, the U.S. Food and Drug Administration (FDA) approved the addition of 0.1 ppm of selenium to all swine diets. In 1982, the FDA approved the addition of 0.3 ppm of selenium to diets for pigs up to 20 kg, because 0.1 ppm of added selenium does not always prevent deficiency signs in weanling pigs (Mahan and Moxon, 1978b; Meyer et al., 1981). The current regulation allows up to 0.3 ppm of selenium in the diet for all pigs (Food and Drug Administration, 1987a,b). As reviewed by Ullrey (1992), concerns about environmental pollution by selenium have led to efforts to reduce the level to 0.1 ppm, but the level of 0.3 ppm has been maintained. The primary biochemical change in selenium deficiency is a decline in glutathione peroxidase activity (Thompson et al., 1976; Young et al., 1976; Fontaine and Valli, 1977). Hence, the level of glutathione peroxidase in the plasma is a reliable index of the selenium status of pigs (Chavez, 1979a,b; Wegger et al., 1980; Adkins and Ewan, 1984). Sudden death is a prominent feature of the seleniumdeficiency syndrome (Ewan et al., 1969; Groce et al., 1971, 1973a,b). The gross necropsy lesions of selenium deficiency are identical to those of vitamin E deficiency. These include massive hepatic necrosis (hepatosis dietetica); edema of the spiral colon, lungs, subcutaneous tissues, and submucosa of the stomach; bilateral paleness and dystrophy of the skeletal muscles (white muscle disease); mottling and dystrophy of the myocardium (mulberry heart disease); impaired reproduction; reduced milk production; and impaired immune response (Eggert et al., 1957; Orstadius et al., 1959; Lindberg and Siren, 1963, 1965; Michel et al., 1969; Trapp et al., 1970; Sharp et al., 1972a,b; Ruth and Van Vleet, 1974; Ullrey, 1974; Fontaine et al., 1977a,b,c; Nielsen et al., 1979; Sheffy and Schultz, 1979; Peplowski et al., 1980; Spallholz, 1980; Larsen and Tollersrud, 1981; Simesen et al., 1982).

When fed to growing swine as sodium selenite, sodium selenate, selenomethionine, or seleniferous corn, selenium does not produce toxicity at levels of less than 5 ppm. In some cases, however, a level of 5 ppm (Mahan and Moxon, 1984) and levels from 7.5 to 10 ppm (Wahlstrom et al., 1955; Trapp et al., 1970; Herigstad et al., 1973; Goehring et al., 1984a,b) have produced toxicity. Signs of toxicity include anorexia, hair loss, fatty infiltration of the liver, degenerative changes in the liver and kidney, edema, occasional separation of hoof and skin at the coronary band (Miller, 1938; Miller and Williams, 1940; Wahlstrom et al., 1955; Orstadius, 1960; Lindberg and Lannek, 1965; Herigstad et al., 1973), and symmetrical, focal areas of vacuolation and neuronal necrosis (Stowe and Herdt, 1992). Dietary arsenicals help to alleviate selenium toxicity (Wahlstrom et al., 1955). Zinc Zinc is a component of many metalloenzymes, including DNA and RNA synthetases and transferases, many digestive enzymes, and is associated with the hormone, insulin. Hence, this element plays an important role in protein, carbohydrate, and lipid metabolism. Many diet-related factors influence the dietary requirement for zinc (Miller et al., 1979), including phytic acid or plant phytates (Oberleas et al., 1962; Oberleas, 1983), calcium (Tucker and Salmon, 1955; Hoekstra et al., 1956; Lewis et al., 1956, 1957a,b; Luecke et al., 1956,1957; Stevenson and Earle, 1956; Bellis and Philp, 1957; Newland et al., 1958; Whiting and Bezeau, 1958; Berry et al., 1961; Hansard and Itoh, 1968; Morgan et al., 1969; Norrdin et al., 1973; Oberleas, 1983), copper (Hoefer et al., 1960; O’Hara et al., 1960; Ritchie et al., 1963; Kirchgessner and Grassman, 1970), cadmium (Pond et al., 1966), cobalt (Hoekstra, 1970), ethylenediamine tetraacetic acid (EDTA) (Owen et al., 1973), histidine (Dahmer et al., 1972a), and protein level and source (Smith et al., 1962; Dahmer et al., 1972b). The zinc requirement of young pigs consuming a casein–glucose diet is low (15 ppm) because such a diet does not contain plant phytates (Smith et al., 1962; Shanklin et al., 1968). For growing pigs fed semipurified diets that contain isolated soybean protein or corn–soybean meal diets (both diets contain significant amounts of phytate) that contain the recommended level of calcium, the zinc requirement is about 50 ppm (Lewis et al., 1956, 1957a,b; Luecke et al., 1956; Stevenson and Earle, 1956; Smith et al., 1958, 1962; Miller et al., 1970). Boars have a higher zinc requirement than gilts; and gilts have a higher requirement than barrows (Liptrap et al., 1970; Miller et al., 1970). The zinc requirement is increased when excessive levels of calcium are fed (Lewis et al., 1956; Forbes, 1960; Hoefer et al., 1960; Pond and Jones, 1964; Pond et al., 1964;

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Oberleas, 1983). The zinc requirement of breeding animals is not well established, but may be higher than for growing pigs due to fetal growth, milk synthesis, tissue repair during uterine involution, and sperm production in boars. A level of 33 ppm of zinc in a corn–soybean meal diet for sows through five parities was adequate for optimal gestation performance, but not for lactation (Hedges et al., 1976). Kirchgessner et al. (1981) estimated the zinc requirement of pregnant sows at 25 ppm in a balance study. A low level of dietary zinc (13 ppm) during the last 4 weeks of pregnancy prolongs the duration of farrowing (Kalinowski and Chavez, 1984). The classic sign of zinc deficiency in growing pigs is hyperkeratinization of the skin, a condition called parakeratosis (Kernkamp and Ferrin, 1953; Tucker and Salmon, 1955). Zinc deficiency reduces the rate and efficiency of growth and levels of serum zinc, alkaline phosphatase, and albumin (Hoekstra et al., 1956, 1967; Luecke et al., 1957; Theuer and Hoekstra, 1966; Miller et al., 1968, 1970; Prasad et al., 1969, 1971; Ku et al., 1970). Gilts fed zincdeficient diets during gestation and lactation produce fewer and smaller pigs, which have reduced serum and tissue zinc levels (Pond and Jones, 1964; Hoekstra et al., 1967; Hill et al., 1983a,b,c). The zinc concentration in the milk from these dams is also reduced (Pond and Jones, 1964). Zinc deficiency retards testicular development of boars and thymic development of young pigs (Miller et al., 1968; Liptrap et al., 1970). Bioavailabilities of zinc from zinc salts vary when these are included in the diet and can be influenced by the type of dietary ingredients used (Miller, 1991). The zinc in zinc sulphate, zinc carbonate, zinc chloride, and zinc metal dust is highly available (100 percent). Bioavailability estimates are expressed as a percentage of a recognized standard and do not refer to percentage absorbed or retained. Absorbed and retained zinc as a percentage of intake is usually much less than 50 percent of the intake. Zinc is less available from zinc oxide (50 to 80 percent) and is poorly available from zinc sulfide (Miller, 1991). Zinc from organic complexes appears to have approximately equal bioavailability to the zinc in zinc sulfate (Hill et al., 1986; Swinkels et al., 1996; Hahn and Baker, 1993; Wedekind et al., 1994; Cheng and Kornegay, 1995; Cheng et al., 1995; Schell and Kornegay, 1996). Zinc from grains and plant protein has low availability (Miller, 1991), but the availability is enhanced by microbial phytase addition to the diet (Kornegay, 1996). Zinc toxicity in growing pigs fed a corn–soybean meal diet supplemented with 2,000 to 4,000 ppm zinc from zinc carbonate was manifested by depression, arthritis, hemorrhage in axillary spaces, gastritis, and death. However, a dietary zinc level of 1,000 ppm was not toxic (Brink et al., 1959). Growing pigs fed 2,000 to 4,000 ppm of zinc from zinc oxide did not show symptoms of zinc toxicity (Cox

57

and Hale, 1962; Hsu et al., 1975; Hill et al., 1983b). However, pigs became lame and unthrifty within 2 months when they were fed a diet containing 1,000 ppm of zinc from zinc lactate (Grimmett et al., 1937). In contrast, pigs fed a diet containing 1,000 ppm of zinc from zinc sulphate for 7 months showed no signs of zinc toxicity (Kulwich et al., 1953). High dietary calcium reduces the severity of zinc toxicity (Hsu et al., 1975). A 5,000-ppm dietary level of zinc as zinc oxide through two parities reduced litter size and pig weight at weaning and caused osteochondrosis in sows (Hill and Miller, 1983; Hill et al., 1983a). Pigs from sows fed high levels of dietary zinc have reduced tissue levels of copper and rapidly develop anemia when fed a low-copper diet (Hill et al., 1983b,c). The toxicity of zinc depends upon the zinc source, dietary level, the duration of feeding, and the levels of other minerals in the diet. A report that reduced postweaning scouring and increased weight gain resulted when the starting diet was supplemented with 3,000 ppm of zinc from zinc oxide for 14 days (Poulsen, 1989) stimulated a great deal of interest in the pharmacological use of zinc. Several recent studies have confirmed this finding and have shown improved weight gain even in the absence of scouring (Kavanagh, 1992; Hahn and Baker, 1993; Carlson et al., 1995; LeMieux et al., 1995; McCully et al., 1995; Smith et al., 1995a,b; Hill et al., 1996). Levels of zinc varied from 2,000 to 6,000 ppm and were fed for up to 5 weeks in some studies. A recent study (Ward et al., 1996) compared zinc oxide and zinc methionine; they reported that supplementing starter diets with 250 ppm zinc from zinc methionine gave equal improvements in performance to 2,000 ppm zinc from zinc oxide. Some studies, however, have failed to observe beneficial effects of pharmacological levels of zinc (Fryer et al., 1992; Tokach et al., 1992; Schell and Kornegay, 1996). A recent large regional study showed that high dietary levels of zinc (3,000 ppm, as zinc oxide) and copper (250 ppm, as copper sulfate) were both efficacious, but were not additive in terms of growth promotion when they were added in combination to diets for weanling pigs (Hill et al., 1996).

REFERENCES Adkins, R. S., and R. C. Ewan. 1984. Effect of selenium on performance, serum selenium concentration and glutathione peroxidase activity in pigs. J. Anim. Sci. 58:346–350. Agricultural Research Council. 1981. The Nutrient Requirements of Pigs: Technical Review. Rev. ed. Slough, England. Commonwealth Agricultural Bureaux. xxii, 307 pp. Alcantara, P. F., L. E. Hanson, and J. D. Smith. 1980. Sodium requirements, balance and tissue composition of growing pigs. J. Anim. Sci. 50:1092–1101. Ammerman, C. B., and S. M. Miller. 1972. Biological availability of minor mineral ions: A review. J. Anim. Sci. 35:681–694.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

58

Nutrient Requirements of Swine

Ammerman, C. B., J. F. Standish, C. E. Holt, R. H. Houser, S. M. Miller, and G. E. Combs. 1974. Ferrous carbonates as sources of iron for weanling pigs and rats. J. Anim. Sci. 38:52–58. Ammerman, C. B., D. H. Baker, and A. J. Lewis, Jr. 1995. Bioavailability of Nutrients for Animals. Amino Acid, Minerals and Vitamins. Academic Press, New York: 441 pp. Amoikon, E. K., J. M. Fernandez, L. L. Southern, D. L. Thompson, Jr., T. L. Ward, and B. M. Olcott. 1995. Effect of chromium tripicolinate on growth, glucose tolerance, insulin sensitivity, plasma metabolites, and growth hormone in pigs. J. Anim. Sci. 73:1123–1130. Anderson, R. A. 1987. Chromium in animal tissues and fluids. Pp. 225–244 in Trace Elements in Human and Animal Nutrition. W. Mertz, ed. Vol. 1. 5th edition. New York: Academic Press, Inc. Andrews, F. N., C. L. Shrewsbury, C. Harper, C. M. Vestal, and L. P. Doyle. 1948. Iodine deficiency in newborn sheep and swine. J. Anim. Sci. 7:298–310. Apgar, G. A., E. T. Kornegay, M. D. Lindemann, and D. R. Notter. 1995. Evaluation of copper sulfate and a copper lysine complex as growth promotants for weanling swine. J. Anim. Sci. 73:2640–2646. Arpi, T., and G. Tollerz. 1965. Iron poisoning in piglets: Autopsy findings in spontaneous and experimental cases. Acta Vet. Scand. 6:360–373. Arrington, L. R., R. N. Taylor, Jr., C. B. Ammerman, and R. L. Shirley. 1965. Effects of excess dietary iodine upon rabbits, hamsters, rats and swine. J. Nutr. 87:394–398. Arthur, John R. 1994. The biochemical functions of selenium: Relationships to thyroid metabolism and antioxidant systems. Pp. 11–20 in Rowett Research Institute Annual Report for 1993, Rowett Research Institute, Bucksburn, Aberdeen, UK. Arthur, S. R., E. T. Kornegay, H. R. Thomas, H. P. Veit, D. R. Notter, and R. A. Barczewski. 1983a. Restricted energy intake and elevated calcium and phosphorus intake for gilts during growth. III. Characterization of feet and limbs and soundness scores of sows during three parities. J. Anim. Sci. 56:876–886. Arthur, S. R., E. T. Kornegay, H. R. Thomas, H. P. Veit, D. R. Notter, K. E. Webb, Jr., and J. L. Baker. 1983b. Restricted energy intake and elevated calcium and phosphorus intake for gilts during growth. IV. Characterization of metacarpal, metatarsal, femur, humerus and turbinate bones of sows during three parities. J. Anim. Sci. 57:1200–1214. Austic, R. E., and C. C. Calvert. 1981. Nutritional interrelationships of electrolytes and amino acids. Fed. Proc. 40:63–67. Baker, D. H. 1977. Sulfur in Nonruminant Nutrition. West Des Moines, Iowa: National Feed Ingredient Association. 123 pp. Baker, D. H. 1991. Bioavailability of minerals and vitamins. Pp. 341–359 in Swine Nutrition, E. R. Miller, D. E. Ullrey, and A. J. Lewis, eds. Boston: Butterworth-Heinemann. Barber, R. S., R. Braude, and K. G. Mitchell. 1955a. Studies on anemia in pigs. 1. The provision of iron by intramuscular injection. Vet. Rec. 67:348–349. Barber, R. S., R. Braude, K. G. Mitchell, and J. Cassidy. 1955b. High copper mineral mixtures for fattening pigs. Chem. Ind. 601–603. Bartley, J. C., E. F. Reber, J. W. Yusken, and H. W. Norton. 1961. Magnesium balance study in pigs three to five weeks of age. J. Anim. Sci. 20:137–141. Bayley, H. S., and R. G. Thomson. 1969. Phosphorus requirements of growing pigs and effect of steam pelleting on phosphorus availability. J. Anim. Sci. 28:484–491. Bayley, H. S., J. Pos, and R. G. Thomson. 1975. Influence of steam pelleting and dietary calcium level on the utilization of phosphorus by the pig. J. Anim. Sci. 40:857–863. Beeson, W. M., F. N. Andrews, H. L. Witz, and T. W. Perry. 1947. The effect of thyroprotein and thiouracil on the growth and fattening of swine. J. Anim. Sci. 6:482. (Abstr.) Bellis, D. B., and J. M. Philp. 1957. Effect of zinc, calcium and phosphorus on the skin and growth of pigs. J. Sci. Food Agric. 8:119–127.

Bengtsson, G., J. Hakkarainen, L. Jonsson, N. Lannek, and P. Lindberg. 1978a. Requirement for selenium (as selenite) and vitamin E (as alphatocopherol) in weaned pigs. I. The effect of varying alpha-tocopherol levels in a selenium-deficient diet on the development of the VESD syndrome. J. Anim. Sci. 47:143–152. Bengtsson, G., J. Hakkarainen, L. Jonsson, N. Lannek, and P. Lindberg. 1978b. Requirement for selenium (as selenite) and vitamin E (as alphatocopherol) in weaned pigs. II. The effect of varying selenium levels in a vitamin E-deficient diet on the development of the VESD syndrome. J. Anim. Sci. 46:153–160. Berry, R. K., M. C. Bell, R. B. Crainger, and R. G. Buescher. 1961. Influence of dietary calcium and zinc on calcium-45, phosphorus 32, and zinc-65 metabolism in swine. J. Anim. Sci. 20:433–439. Bertram, M. J., T. S. Stahly, R. C. Ewan. 1994. Impact of lean growth genotype and dietary phosphorus regimen on rate and efficiency of growth and carcass characteristics of pigs. J. Anim. Sci. 72(Suppl. 2):68 (Abstr.). Biehl, R. R., and D. H. Baker. 1996. Efficacy of supplemental 1 ahydroxycholecalciferol and microbial phytase for young pigs fed phosphorus- or amino acid-deficient corn-soybean meal diets. J. Anim. Sci. 74:2960–2966. Biehl, R. R., J. L. Emmert, and D. H. Baker. 1997. Iron bioavailability in soybean meal as affected by supplemental phytase and 1 a-hydroxycholecalciferol. Poultry Sci. 76:1424–1427. Blair, R., and D. Benzie. 1964. The effect of level of dietary calcium and phosphorus on skeletal development in the young pig to 25-lb liveweight. Br. J. Nutr. 18:91–101. Bohstedt, G., and R. H. Grummer. 1954. Salt poisoning of pigs. J. Anim. Sci. 13:933–939. Boleman, S. L., S. J. Boleman, T. D. Bidner, L. L. Southern, T. L. Ward, J. E. Pontif, and M. M. Pike. 1995. Effect of chromium picolinate on growth, body composition, and tissue accretion in pigs. J. Anim. Sci. 73:2033–2042. Bonnette, E. D., E. T. Kornegay, M. D. Lindemann, and D. R. Notter. 1990. Influence of two supplemental vitamin E levels and weaning age on performance, humoral antibody production, and serum cortisol levels of pigs. J. Anim. Sci. 68:1346–1353. Bowland, J. P., R. Braude, A. G. Chamberlain, R. F. Glascock, and K. G. Mitchell. 1961. The absorption, distribution and excretion of labeled copper in young pigs given different quantities, as sulphate or sulphide, orally or intravenously. Br. J. Nutr. 15:59–72. Boyd, R. D., D. Hall, and J. F. Wu. 1983. Plasma alkaline phosphatase as a criterion for determining biological availability of phosphorus for swine. J. Anim. Sci. 57:396–401. Brady, P. S., P. K. Ku, D. E. Ullrey, and E. R. Miller. 1978. Evaluation of an amino acid-iron chelate hematinic for the baby pig. J. Anim. Sci. 47:1135–1140. Braude, R. 1967. Copper as a stimulant in pig feeding (Cuprum pro pecunia). World Rev. Anim. Prod. 3:69–82. Braude, R. 1975. Copper as a performance promoter in pigs. Pp. 79–97 in Proc. Copper in Farming Symp. London: Copper Development Association. Braude, R., and E. Cotchin. 1949. Thiourea and methylthiouracil as supplements in rations of fattening pigs. Br. J. Nutr. 3:171–186. Braude, R., A. G. Chamberlain, M. Kotarbinska, and K. G. Mitchell. 1962. The metabolism of iron in piglets given labeled iron either orally or by injection. Br. J. Nutr. 16:427–449. Brennan, J. J., and F. X. Aherne. 1984. Effect of calcium and phosphorus levels in the diet on the incidence of leg weakness in swine. Agric. Forestry Bull. Special Issue. Pp. 8–10. Brink, M. F., D. E. Becker, S. W. Terrill, and A. H. Jensen. 1959. Zinc toxicity in the weanling pig. J. Anim. Sci. 18:836–842. Bunch, R. J., J. T. McCall, V. C. Speer, and V. W. Hays. 1965. Copper supplementation for weanling pigs. J. Anim. Sci. 24:995–1000.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Burnell, T. W., G. L. Cromwell, and T. S. Stahly. 1988. Effects of particle size on the availability of phosphorus in defluorinated phosphate for pigs. J. Anim. Sci. 67(Suppl. 1):140 (Abstr.). Calabotta, D. F., E. T. Kornegay, H. R. Thomas, J. W. Knight, D. R. Notter, and H. P. Veit. 1982. Restricted energy intake and elevated calcium and phosphorus intake for gilts during growth. I. Feedlot performance and foot and leg measurements and scores during growth. J. Anim. Sci. 54:565–575. Calvert, C. C., R. J. Besecker, M. P. Plumlee, T. R. Cline, and D. M. Forsyth. 1978. Apparent digestibility of phosphorus in barley and corn for growing swine. J. Anim. Sci. 47:420–426. Campbell, E. A. 1961. Iron poisoning in the young pig. Aust. Vet. J. 37:78–83. Campbell, R. G. 1996. The effects of chromium picolinate on the fertility and fecundity of sows under commercial conditions. Pp. 33–38 in Sixteenth Annual Feed Ingredient Conference, Prince Agri. Products, Inc., Quincy, IL, Aug. 19–20. Carlson, M. S., G. M. Hill, J. E. Link, G. A. McCully, D. W. Rozeboom, and R. L. Weavers. 1995. Impact of zinc oxide and copper sulfate supplementation on the newly weaned pig. J. Anim. Sci. 73(Suppl. 1):72 (Abstr.). Carnes, W. H., C. S. Shields, C. E. Cartwright, and M. M. Winthrop. 1961. Vascular lesions in copper-deficient swine. Fed. Proc. 20:118. (Abstr.) Carson, T. L. 1986. Toxic chemicals, plants, metals and mycotoxins. Pp. 688–701 in Diseases of Swine, 6th Ed., A. D. Leman, B. Straw, R. D. Glock. W. L. Mengeling, R. H. C. Penny, and E. Scholl, eds. Ames: Iowa State University Press. Carter, J. H., Jr., R. F. Miller, and C. C. Brooks. 1959. The effect of copper and calcium levels on the performance of growing swine. J. Anim. Sci. 18:1502 (Abstr.). Carter, S. D., and G. L. Cromwell. 1998a. Influence of porcine somatotropin on the phosphorus requirement of finishing pigs. I. Performance and bone characteristics. J. Anim. Sci. 76:584–595. Carter, S. D., and G. L. Cromwell. 1998b. Influence of porcine somatotropin on the phosphorus requirement of finishing pigs. II. Carcass characteristics, tissue accretion rates, and chemical composition of the ham. J. Anim. Sci. 76:596–605. Chaney, C. H., and C. E. Barnhart. 1963. Effect of iron supplementation of sow rations on the prevention of baby pig anemia. J. Nutr. 81:187–192. Chapman, H. L., Jr., J. Kastelic, C. C. Ashton, and D. V. Catron. 1955. A comparison of phosphorus from different sources for growing and finishing swine. J. Anim. Sci. 14:1073–1085. Chapman, H. L., Jr., J. Kastelic, G. C. Ashton, P. G. Homeyer, C. Y. Roberts, D. V. Catron, V. W. Hays, and V. C. Speer. 1962. Calcium and phosphorus requirements for growing-finishing swine. J. Anim. Sci. 21:112–118. Chavez, E. R. 1979a. Effects of dietary selenium depletion and repletion on plasma glutathione peroxidase activity and selenium concentration in blood and body tissue of growing pigs. Can. J. Anim. Sci. 59:761–771. Chavez, E. R. 1979b. Effects of dietary selenium on glutathione peroxidase activity in piglets. Can. J. Anim. Sci. 59:67–75. Chavez, E. R. 1985. Nutritional significance of selenium supplementation in a semi-purified diet fed during gestation and lactation to first-litter gilts and their piglets. Can. J. Anim. Sci. 64:497–506. Cheng, J., and E. T. Kornegay. 1995. Comparison of zinc sulfate and a zinc complex as zinc source for young pigs fed lysine deficient and adequate diets. J. Anim. Sci. 73(Suppl. 1):172 (Abstr.). Cheng, J., T. C. Schell, and E. T. Kornegay. 1995. Influence of dietary lysine on the utilization of zinc from zinc sulfate and a zinc lysine complex by young pigs. J. Anim. Sci. 73(Suppl. 1):17 (Abstr.). Christianson, S. L., E. R. Peo, Jr., and A. J. Lewis. 1989. Effects of dietary manganese levels on reproductive performance of sows. J. Anim. Sci. 67(Suppl. 1):251 (Abstr.).

59

Christianson, S. L., E. R. Peo, Jr., A. J. Lewis, and M. A. Giesemann. 1990. Influence of dietary manganese levels on reproduction, serum cholesterol and milk manganese concentration of sows. J. Anim. Sci. 68(Suppl. 1):368 (Abstr.). Chung, A. S., W. C. Hoekstra, and R. H. Grummer. 1976. Supplemental cobalt or nickel for zinc-deficient G.F. pigs. J. Anim. Sci. 42:1352 (Abstr.). Clancy, S. P., P. M. Clarkson, M. E. DeCheke, K. Nosaka, P. S. Freedson, J. J. Cunningham, and B. Valentine. 1994. Effects of chromium picolinate supplementation on body composition, strength, and urinary chromium loss in football players. Int. J. Sports Nutr. 4:142–153. Clawson, A. J., and W. D. Armstrong. 1981. Ammonium polyphosphate as a source of phosphorus and nonprotein nitrogen for monogastrics. J. Anim. Sci. 52:1–7. Coalson, J. A., C. V. Maxwell, J. C. Hillier, R. D. Washam, and E. C. Nelson. 1972. Calcium and phosphorus requirements of young pigs reared under controlled environmental conditions. J. Anim. Sci. 35: 1194–1200. Coalson, J. A., C. V. Maxwell, J. C. Hillier, and E. C. Nelson. 1974. Calcium requirement of the cesarean derived colostrum-free pig from 3 through 9 weeks of age. J. Anim. Sci. 38:772–777. Coffey, R. D., and G. L. Cromwell. 1993. Evaluation of the biological availability of phosphorus in various feed ingredients for growing pigs. J. Anim. Sci. 71(Suppl. 1):66 (Abstr.). Coffey, R. D., G. L. Cromwell, and H. J. Monegue. 1994. Efficacy of a copper-lysine complex as a growth promotant for weanling pigs. J. Anim. Sci. 72:2880–2886. Coffey, R. D., K. W. Mooney, G. L. Cromwell, and D. K. Aaron. 1994. Biological availability of phosphorus in defluorinated phosphates with different phosphorus solubilities in neutral ammonium citrate for chicks and pigs. J. Anim. Sci. 72:2653–2660. Combs, C. E., and H. D. Wallace. 1962. Growth and digestibility studies with young pigs fed various levels and sources of calcium. J. Anim. Sci. 21:734–737. Combs, G. E., J. M. Vandepopuliere, H. D. Wallace, and M. Koger. 1962. Phosphorus requirement of young pigs. J. Anim. Sci. 21:3–8. Combs, G. E., T. H. Berry, H. D. Wallace, and R. C. Crum, Jr. 1966. Levels and sources of vitamin D for pigs fed diets containing varying quantities of calcium. J. Anim. Sci. 25:827–830. Combs, N. R., and E. R. Miller. 1985. Determination of potassium availability in K2CO3, KHCO3, corn and soybean meal for the young pig. J. Anim. Sci. 60:715–719. Combs, N. R., E. R. Miller, and P. K. Ku. 1985. Development of an assay to determine the bioavailability of potassium in feedstuffs for the young pig. J. Anim. Sci. 60:709–714. Combs, N. R., E. T. Kornegay, M. D. Lindemann, and D. R. Notter. 1991a. Calcium and phosphorus requirement of swine from weaning to market: I. Development of response curves for performance. J. Anim. Sci. 69:673–681. Combs, N. R., E. T. Kornegay, M. D. Lindemann, D. R. Notter, J. W. Wilson, and J. P. Mason. 1991b. Calcium and phosphorus requirement of swine from weaning to market: II. Development of response curves for bone criteria and comparison of bending and shear bone testing. J. Anim. Sci. 69:682–693. Corley, J. R., D. H. Baker, and R. A. Easter. 1980. Biological availability of phosphorus in rice bran and wheat bran as affected by pelleting. J. Anim. Sci. 50:286–292. Coulter, D. B., and M. J. Swenson. 1970. Effects of potassium intoxication on porcine electrocardiograms. Am. J. Vet. Res. 31:2001–2011. Cox, D. H., and O. M. Hale. 1962. Liver iron depletion without copper loss in swine for excess zinc. J. Nutr. 77:225–228. Cox, J. L., D. E. Becker, and A. H. Jensen. 1966. Electrocardiographic evaluation of potassium deficiency in young swine. J. Anim. Sci. 25:203–206.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

60

Nutrient Requirements of Swine

Crenshaw, T. D., E. R. Peo, Jr., A. J. Lewis, B. D. Moser, and D. Olson. 1981. Influence of age, sex and calcium and phosphorus levels on the mechanical properties of various bones in swine. J. Anim. Sci. 52:1319–1329. Cromwell, G. L. 1979. Availability of phosphorus in feedstuffs for swine. Proc. Distill. Feed. Conf. 34:40–52. Cromwell, G. L. 1985. Phosphorus requirements of swine. Pp. 48–65 in Proc. 8th Annual International Minerals Conference. Mundelein, IL: International Minerals and Chemical Corp. Cromwell, G. L. 1992. The biological availability of phosphorus in feedstuffs for pigs. Pig News and Information 13:75N. Cromwell, G. L. 1997. Copper as a nutrient for animals. Pp. 177–202 in Handbook of Copper Compounds and Applications, H. W. Richardson, Ed. New York: Marcel Dekker, Inc. Cromwell, G. L., V. W. Hays, C. H. Chaney, and J. R. Overfield. 1970. Effects of dietary phosphorus and calcium level on performance, bone mineralization and carcass characteristics of swine. J. Anim. Sci. 30:519–525. Cromwell, G. L., V. W. Hays, and J. R. Overfield. 1972a. Effects of phosphorus levels in corn-, milo- and wheat-based diets on performance and bone strength of pigs. J. Anim. Sci. 35:1103(Abstr.). Cromwell, G. L., V. W. Hays, C. W. Scherer, and J. R. Overfield. 1972b. Effects of dietary calcium and phosphorus on performance and carcass, metacarpal and turbinate characteristics of swine. J. Anim. Sci. 34:746–751. Cromwell, G. L., V. W. Hays, and J. R. Overfield. 1974. Effects of phosphorus levels in corn, wheat and barley diets on performance and bone strength of swine. J. Anim. Sci. 39: 180(Abstr.). Cromwell, G. L., D. T. H. Sihombing, and V. W. Hays. 1975. Effects of iodine level on performance and thyroid traits of growing pigs. J. Anim. Sci. 41:813–818. Cromwell, G. L., V. W. Hays, J. R. Overfield, and J. L. Krug. 1976. Meat and bone meal as a source of phosphorus for growing swine. J. Anim. Sci. 42:1350(Abstr.). Cromwell, G. L., V. W. Hays, and T. L. Clark. 1978. Effects of copper sulfate, copper sulfide and sodium sulfide on performance and copper stores of pigs. J. Anim. Sci. 46:692–698. Cromwell, G. L., T. S. Stahly, and J. R. Overfield. 1979. Effect of dietary phosphorus level on growth, carcass and bone traits of boars. J. Anim. Sci. 49(Suppl. 1):101 (Abstr.). Cromwell, G. L., T. S. Stahly, and H. J. Monegue. 1981a. Effects of sodium and chloride on performance of pigs J. Anim. Sci. 53(Suppl. 1):237 (Abstr.). Cromwell, G. L., T. S. Stahly, and W. D. Williams. 1981b. Efficacy of copper as a growth promotant and its interrelation with sulfur and antibiotics for swine. Distill. Feed Conf. Proc. Distill. Feed Res. Counc. 36:6. Cromwell, G. L., T. S. Stahly, and H. J. Monegue. 1983. Bioavailability of the calcium and phosphorus in dehydrated alfalfa meal for growing pigs. J. Anim. Sci. 57(Suppl. 1):242 (Abstr.). Cromwell, G. L., T. S. Stahly, and H. J. Monegue. 1985. Bioavailability of the phosphorus in wheat for the pig. J. Anim. Sci.61(Suppl. 1):320 (Abstr.). Cromwell, G. L., T. S. Stahly, and T. V. Burnell. 1987. Source and particle size of phosphorus supplements on bioavailability of phosphorus. J. Anim. Sci. 65(Suppl. 1):134 (Abstr.). Cromwell, G. L., D. D. Hall, G. E. Combs, O. M. Hale, D. L. Handlin, J. P. Hitchcock, D. A. Knabe, E. T. Kornegay, M. D. Lindemann, C. V. Maxwell, and T. J. Prince. 1989a. Effects of dietary salt level during gestation and lactation on reproductive performance of sows: A cooperative study. J. Anim. Sci. 67:374–385. Cromwell, G. L., H. J. Monegue, and T. S. Stahly. 1989b. Effects of source and level of copper on performance and liver copper stores in weanling pigs. J. Anim. Sci. 67:2996–3002.

Cromwell, G. L., H. J. Monegue, and T. S. Stahly. 1993a. Long-term effects of feeding a high copper diet to sows during gestation and lactation. J. Anim. Sci. 71:2996–3002. Cromwell, G. L., T. S. Stahly, R. D. Coffey, H. J. Monegue, and J. H. Randolph. 1993b. Efficacy of phytase in improving the bioavailability of phosphorus in soybean meal and corn-soybean meal diets for pigs. J. Anim. Sci. 71:1831–1840. Cromwell, G. L., M. D. Lindemann, H. J. Monegue, D. D. Hall, and D. E. Orr, Jr. 1998. Tribasic copper chloride and copper sulfate as copper sources for weanling pigs. J. Anim. Sci. 76:118–123. Cromwell, G. L., R. D. Coffey, G. R. Parker, H. J. Monegue, and J. H. Randolph. 1995. Efficacy of a recombinant-derived phytase in improving the bioavailability of phosphorus in corn-soybean meal diets for pigs. J. Anim. Sci. 73:2000–2008. Cromwell, G. L., J. L. Pierce, T. E. Sauber, D. W. Rice, D. S. Etrl, and V. Raboy. 1998. Bioavailability of phosphorus in low-phytic acid corn for growing pigs. J. Anim. Sci. 76(Suppl. 2): (Midwestern Section Abstr. #115). Dahmer, E. J., B. W. Coleman, R. H. Grummer, and W. G. Hoekstra. 1972a. Alleviation of parakeratosis in zinc-deficient swine by high levels of dietary histidine. J. Anim. Sci. 35:1181–1189. Dahmer, E. J., R. H. Grummer, and W. G. Hoekstra. 1972b. Prevention of zinc deficiency in swine by feeding blood meal. J. Anim. Sci. 35:1176–1180. Devilat, J., and A. Skoknic. 1971. Feeding high levels of rapeseed meal to pregnant gilts. Can. J. Anim. Sci. 51:715–719. Diehl, J. S., D. C. Mahan, and A. L. Moxon. 1975. Effects of single intramuscular injections of selenium at various levels to young swine. J. Anim. Sci. 40:844–850. Doige, C. E., B. D. Owen, and J. H. L. Mills. 1975. Influence of calcium and phosphorus on growth and skeletal development of growing swine. Can. J. Anim. Sci. 55:147–164. Doornenbal, H. 1975. Tissue selenium content of the growing pig. Can. J. Anim. Sci. 55:325–330. Dove, C. R. 1993a. The effect of adding copper and various fat sources to the diets of weanling swine on growth performance and serum fatty acid profiles. J. Anim. Sci. 71:2187–2192. Dove, C. R. 1993b. The effect of copper supplementation during lactation on sow and pig performance and the subsequent nursery performance of pigs. J. Anim. Sci. 71(Suppl. 1):173 (Abstr.). Dove, C. R. 1995. The effect of copper level on nutrient utilization of weanling pigs. J. Anim. Sci. 73:166–171. Dove, C. R. and K. D. Haydon. 1992. The effect of copper and fat addition to the diets of weanling swine on growth performance and serum fatty acids. J. Anim. Sci. 70:805–810. Ducsay, C. A., W. C. Buhi, F. W. Bazer, R. M. Roberts, and C. E. Combs. 1984. Role of uteroferrin in placental iron transport: Effect of maternal iron treatment on fetal iron and uteroferrin content and neonatal hemoglobin. J. Anim. Sci. 59:1303–1308. Edmonds, M. S., O. A. Izquierdo, and D. H. Baker. 1985. Feed additive studies with newly weaned pigs: Efficacy of supplemental copper, antibiotics and organic acids. J. Anim. Sci. 60:462–469. Eeckhout, W., M. de Paepe, N. Warnants, H. Bekaert. 1995. An estimation of the minimal P requirements for growing-finishing pigs, as influenced by the Ca level of the diet. Anim. Feed Sci. Tech. 52:29–40. Eggert, R. O., D. Patterson, W. T. Akers, and E. L. R. Stokstad. 1957. The role of vitamin E and selenium in the nutrition of the pig. J. Anim. Sci. 16:1037(Abstr.). Elvehjem, C. A., and E. B. Hart. 1932. The necessity of copper as a supplement to iron for hemoglobin formation in the pig. J. Biol. Chem. 95:363–370. Evock-Clover, C. M., M. M. Polansky, R. A. Anderson, and N. C. Steele. 1993. Dietary chromium supplementation with or without somatotropin treatment alters serum hormones and metabolites in growing pigs without affecting growth performance. J. Nutr. 123:1504–1512.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Ewan, R.C. 1971. Effect of vitamin E and selenium on tissue composition of young pigs. J. Anim. Sci. 32:883–887. Ewan, R. C., M. E. Wastell, E. J. Bicknell, and V. C. Speer. 1969. Performance and deficiency symptoms of young pigs fed diets low in vitamin E and selenium. J. Anim. Sci. 29:912–915. Fammatre, C. A., D. C. Mahan, A. W. Fetter, A. P. Crifo, Jr., and J. K. Judy. 1977. Effects of dietary protein, calcium and phosphorus levels for growing and finishing swine. J. Anim. Sci. 44:65–71. Farries, F. E. 1958. Kali-Briefe (Tierzucht) no. 3. 1981. P. 290 in The Nutrient Requirements of Pigs. Agricultural Research Council. Slough, England: Commonwealth Agricultural Bureaux. Follis, R. H., J. A. Bush, G. E. Cartwright, and M. W. Wintrobe. 1955. Studies on copper metabolism. XVIII. Skeletal changes associated with copper deficiency in swine. Bull. Johns Hopkins Hosp. 97:405. Fontaine, M., and V. E. O. Valli. 1977. Studies on vitamin E and selenium deficiency in young pigs. II. The hydrogen peroxide hemolysis test and the measure of red cell lipid peroxides as indices of vitamin E and selenium status. Can. J. Comp. Med. 41:52–56. Fontaine, M., V. E. O. Valli, and L. G. Young. 1977a. Studies on vitamin E and selenium deficiency in young pigs. IV. Effect on coagulation system. Can. J. Comp. Med. 41:64–76. Fontaine, M., V. E. O. Valli, and L. G. Young. 1977b. Studies on vitamin E and selenium deficiency in young pigs. III. Effect on kinetics of erythrocyte production and destruction. Can. J. Comp. Med. 41:57–63. Fontaine, M., V. E. O. Valli, L. G. Young, and J. H. Lumsden. 1977c. Studies on vitamin E and selenium deficiency in young pigs. I. Hematological and biochemical changes. Can. J. Comp. Med. 41:41–51. Food and Drug Administration. 1987a. Food additives permitted in feed and drinking water of animals: Selenium. Federal Register 52:10887 (Monday, April 6). Food and Drug Administration. 1987b. Food additives permitted in feed and drinking water of animals: Selenium; Correction. Federal Register 52:21001 (Thursday, June 4). Forbes, R. M. 1960. Nutritional interactions in zinc and calcium. Fed. Proc. 19:643–647. Friend, D. W., and M. S. Wolynetz. 1981. Self-selection of salt by gilts during pregnancy and lactation. Can. J. Anim. Sci. 61:429–438. Froseth, J. A., D. C. Honeyfield, and R. J. Barke. 1982a. Dietary sodium and chloride levels for young pigs. J. Anim. Sci. 55(Suppl. 1):271 (Abstr.). Froseth, J. A., P. K. Ku, and E. R. Miller. 1982b. Addition of potassium carbonate to swine diets containing varying levels of lysine. J. Anim. Sci. 55(Suppl. 1):97 (Abstr.). Froseth, J. A., E. R. Miller, P. K. Ku, and W. G. Bergen. 1982c. Effects and interactions of dietary lysine, potassium and sodium levels for young pigs. J. Anim. Sci. 55(Suppl. 1):271 (Abstr.). Fryer, A. J., P. K. Ku, E. R. Miller, and D. E. Ullrey. 1992. Effect of elevated dietary zinc on growth performance of weanling swine. J. Anim. Sci. 70(Suppl. 1):62 (Abstr.). Fuller, R., L. G. M. Newland, C. A. E. Briggs, R. Braude, and K. G. Mitchell. 1960. The normal intestinal flora of the pigs. IV. The effect of dietary supplements of penicillin, chlortetracyline or copper on the faecal flora. J. Appl. Bacterial. 23:195–205. Furugouri, K. 1972. Effect of elevated dietary levels of iron on iron store in liver, some blood constituents and phosphorus deficiency in young swine. J. Anim. Sci. 34:573–577. Furugouri, K., and A. Kawabata. 1975. Iron absorption in nursing piglets. J. Anim. Sci. 41:1348–1354. Furugouri, K., and A. Kawabata. 1976. Iron absorption by neonatal pig intestine in vivo. J. Anim. Sci. 42:1460–1464. Furugouri, K., and A. Kawabata. 1979. Iron absorptive function of neonatal pig intestine. J. Anim. Sci. 49:715–723. Gamble, C. T., S. L. Hansard, B. R. Moss, D. J. Davis, and E. R. Lidvall. 1971. Manganese utilization and placental transfer in the gravid gilt. J. Anim. Sci. 32:84–87.

61

Gargas, M. L., R. I. Norton, D. J. Paustenbach, and B. L. Finley. 1991. Urinary excretion of chromium by humans following ingestion of chromium picolinate: Implications for biomonitoring. Drug Metab. Dispos. 22:522–529. Gleed, P. T., and B. F. Sansom. 1982. Ingestion of iron in sow’s faeces by piglets reared in farrowing crates with slotted floors. Br. J. Nutr. 47:113–117. Giesemann, M. A., A. J. Lewis, M. P. Akhter, and P. S. Miller. 1992a. Effect of the reproductive cycle on bone strength of young and mature sows. J. Anim. Sci. 70(Suppl. 1):230 (Abstr.). Giesemann, M. A., A. J. Lewis, and P. S. Miller. 1992b. Calcium and phosphorus balance of gilts and mature sows during gestation and lactation. J. Anim. Sci. 70(Suppl. 1):69 (Abstr.). Glienke, L. R., and R. C. Ewan. 1977. Selenium deficiency in the young pig. J. Anim. Sci. 45: 1334–1340. Goehring, T. B., I. S. Palmer, O. E. Olson, C. W. Libal, and R. C. Wahlstrom. 1984a. Effects of seleniferous grains and inorganic selenium on tissue and blood composition and growth performance of rats and swine. J. Anim. Sci. 59:725–732. Goehring, T. B., I. S. Palmer, O. E. Olson, C. W. Libal, and R. C. Wahlstrom. 1984b. Toxic effects of selenium on growing swine fed corn-soybean meal diets. J. Anim. Sci. 59:733–737. Golz, D. I., and T. D. Crenshaw. 1990. Interrelationships of dietary sodium, potassium and chloride on growth in young swine. J. Anim. Sci. 68:2736–2747. Golz, D. I., and T. D. Crenshaw. 1991. The effect of dietary potassium and chloride on cation-anion balance in swine. J. Anim. Sci. 69:2504–2515. Grandhi, R. R., and J. H. Strain. 1983. Dietary calcium-phosphorus levels for growth and reproduction in gilts and sows. Can. J. Anim. Sci. 63:443–454. Grant, C. A., B. Thafvelin, and R. Christell. 1961. Retention of selenium by pig tissues. Acta Pharmacol. Toxicol. 18:285–297. Grimmett, R. E. R., I. G. McIntosh, E. M. Wall, and C. S. M. Hopkirk. 1937. Chromium zinc poisoning of pigs. Results of experimental feeding of pure zinc lactate. New Zealand J. Agric. 54:216–223. Groce, A. W., E. R. Miller, K. K. Keahey, D. E. Ullrey, and D. J. Ellis. 1971. Selenium supplementation of practical diets for growing-finishing swine. J. Anim. Sci. 32:905–911. Groce, A. W., E. R. Miller, D. E. Ullrey, P. K. Ku, K. K. Keahey, and D. J. Ellis. 1973a. Selenium requirements in corn-soy diets for growingfinishing swine. J. Anim. Sci. 37:948–956. Groce, A. W., E. R. Miller, J. P. Hitchcock, D. E. Ullrey, and W. T. Magee. 1973b. Selenium balance in the pig as affected by selenium source and vitamin E. J. Anim. Sci. 37:942–947. Grummer, R. H., O. G. Bentley, P. H. Phillips, and G. Bohstedt. 1950. The role of manganese in growth, reproduction and lactation of swine. J. Anim. Sci. 9:170–175. Hagsten, I., and T. W. Perry. 1976a. Evaluation of dietary salt levels for swine. I. Effect on gain, water consumption and efficiency of feed conversion. J. Anim. Sci. 42:1187–1190. Hagsten, I., and T. W. Perry. 1976b. Evaluation of dietary salt levels for swine. II. Effect on blood and excretory patterns. J. Anim. Sci. 42:1191–1195. Hagsten, I., T. R. Cline, T. W. Perry, and M. P. Plumlee. 1976. Salt supplementation of corn-soy diets for swine. J. Anim. Sci. 42:12–15. Hahn, J. D., and D. H. Baker. 1993. Growth and plasma zinc responses of young pigs fed pharmacologic levels of zinc. J. Anim. Sci. 71:3020–3024. Hakkarainen, J., P. Lindberg, G. Bengtsson, L. Jonsson, and N. Lannek. 1978. Requirement for selenium (as selenite) and vitamin E (as alphatocopherol) in weaned pigs. III. The effect on the development of the VESD syndrome of varying selenium levels with a low- tocopherol diet. J. Anim. Sci. 46:1001–1008. Hall, D. D., G. L. Cromwell, and T. S. Stahly. 1991. Effects of dietary calcium, phosphorus, calcium:phosphorus ratio and vitamin K on per-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

62

Nutrient Requirements of Swine

formance, bone strength and blood clotting status of pigs. J. Anim. Sci. 69:646–655. Hall, V. L., R. C. Ewan, and M. J. Wannemuehler. 1993. Effect of zinc deficiency and zinc source on performance and immune response in young pigs. J. Anim. Sci. 71(Suppl. 1):173 (Abstr.). Hancock, J. E., E. R. Peo, Jr., A. J. Lewis, J. D. Crenshaw, and B. D. Moser. 1986. Vitamin D toxicity in young pigs. J. Anim. Sci. 63(Suppl. 1):268 (Abstr.). Hansard, S. L., and H. Itoh. 1968. Influence of limited dietary calcium upon zinc absorption, placental transfer and utilization by swine. J. Nutr. 95:23–30. Hansen, B. C., A. J. Lewis, and E. R. Peo, Jr. 1987. Bone traits of growing boars, barrows and gilts fed different levels of dietary protein and available phosphorus. J. Anim. Sci 65(Suppl. 1):126 (Abstr.). Harmon, B. G., D. E. Becker, and A. H. Jensen. 1967. Efficacy of ferric ammonium citrate in preventing anemia in young swine. J. Anim. Sci. 26:1051–1053. Harmon, B. G., D. E. Hoge, A. H. Jensen, and D. H. Baker. 1969. Efficacy of ferrous carbonate as a hematinic for young swine. J. Anim. Sci. 29:706–710. Harmon, B. C., S. G. Cornelius, J. Totsch, D. H. Baker, and A. H. Jensen. 1974a. Oral iron dextran and iron from steel slats as hematinics for swine. J. Anim. Sci. 39:699–702. Harmon, B. G., C. T. Liu, S. G. Cornelius, J. E. Pettigrew, D. H. Baker, and A. H. Jensen. 1974b. Efficacy of different phosphorus supplements for sows during gestation and lactation. J. Anim. Sci. 39:1117–1123. Harmon, B. G., C. T. Liu, A. H. Jensen, and D. H. Baker. 1975. Phosphorus requirements of sows during gestation and lactation. J. Anim. Sci. 40:660–664. Harmon, B. G., C. T. Liu, A. H. Jensen, and D. H. Baker. 1976. Dietary magnesium levels for sows during gestation and lactation. J. Anim. Sci. 42:860–865. Harper, A. F., M. D. Lindemann, and E. T. Kornegay. 1995. Effect of supplemental dietary chromium on growth performance of weanling swine. J. Anim. Sci. 73(Suppl. 1):194 (Abstr.) Harris, J. E., S. D. Crow, and M. D. Newcomb. 1995. Effect of chromium picolinate on growth performance and carcass characteristics on pigs fed adequate and low protein diets. J. Anim. Sci. 73(Suppl. 1):194 (Abstr.). Harrold, R. L. 1981. Digestible energy and available phosphorus content of sunflower seed products. J. Anim. Sci. 53(Suppl. 1):516 (Abstr.). Hart, E. B., and H. Steenbock. 1918. Hairless pigs: The cause and remedy. Wis. Agric. Exp. Stn. Bull. 297:1–11. Hart, E. B., C. A. Elvehjem, H. Steenbock, A. R. Kemmerer, G. Bohstedt, and J. M. Fargo. 1930. A study of the anemia of young pigs and its prevention. J. Nutr. 2:277–294. Haydon, K. K., J. W. West, and M. N. McCarter. 1993. Effect of dietary electrolyte balance on performance and blood parameters of growingfinishing swine fed in high ambient temperatures. J. Anim. Sci.68:2400–2406. Hayes, S. H., G. L. Cromwell, T. S. Stahly, and T. H. Johnson. 1979. Availability of phosphorus in corn, wheat and barley for the chick. J. Anim. Sci. 49:992–992. Hays, V. W. 1976. Phosphorus in Swine Nutrition. West Des Moines, Iowa: National Feed Ingredients Association. Hedges, J. D., and E. T. Kornegay. 1973. Interrelationship of dietary copper and iron as measured by blood parameters, tissue stores and feedlot performance of swine. J. Anim. Sci. 37:1147–1154. Hedges, J. D., E. T. Kornegay, and H. R. Thomas. 1976. Comparison of dietary zinc levels for reproducing sows and the effect of dietary zinc and calcium on the subsequent performance of their progeny. J. Anim. Sci. 43:453–463. Herigstad, R. R., C. K. Whitehair, and O. E. Olson. 1973. Inorganic and organic selenium toxicosis in young swine: Comparison of pathologic

changes with those in swine with vitamin E-selenium deficiency. Am. J. Vet. Res. 34:1227–1238. Hew, V. F., G. L. Cromwell, and T. S. Stahly. 1982. The bioavailability of phosphorus in some tropical feedstuffs for pigs. J. Anim. Sci. 55(Suppl. 1):277 (Abstr.). Hickman, D. S., D. C. Mahan, and J. H. Cline. 1983. Dietary calcium and phosphorus for developing boars. J. Anim. Sci. 56:431–437. Hill, D. A., E. R. Peo, Jr., A. J. Lewis, and J. D. Crenshaw. 1986. Zincamino acid complexes for swine. J. Anim. Sci. 63:121–130. Hill, G. M., and E. R. Miller. 1983. Effect of dietary zinc levels on the growth and development of the gilt. J. Anim. Sci. 57:106–113. Hill, G. M., P. K. Ku, E. R. Miller, D. E. Ullrey, T. A. Losty, and B. L. O’Dell. 1983a. A copper deficiency in neonatal pigs induced by a high zinc maternal diet. J. Nutr. 113:867–872. Hill, G. M., E. R. Miller, and H. D. Stowe. 1983b. Effect of dietary zinc levels on health and productivity of gilts and sows through two parities. J. Anim. Sci. 57:114–122. Hill, G. M., E. R. Miller, P. A. Whetter, and D. E. Ullrey. 1983c. Concentrations of minerals in tissues of pigs from dams fed different levels of dietary zinc. J. Anim. Sci. 57:130–138. Hill, G. M., G. L. Cromwell, T. D. Crenshaw, R. C. Ewan, D. A. Knabe, A. J. Lewis, D. C. Mahan, G. C. Shurson, L. L. Southern, and T. L. Veum. 1996. Impact of pharmacological intakes of zinc and (or) copper on performance of weanling pigs. J. Anim. Sci. 74(Suppl. 1):181 (Abstr.). Hitchcock, J. P., P. K. Ku, and E. R. Miller. 1974. Factors influencing iron utilization by the baby pig. Pp. 598–600 in Trace Element Metabolism in Animals, Volume a, W. G. Hoekstra, J. W. Suttie, H. E. Ganther, and W. Mertz, eds. Baltimore: University Park Press. Hoefer, J. A., E. R. Miller, D. E. Ullrey, H. D. Ritchie, and R. W. Luecke. 1960. Interrelationships between calcium, zinc, iron and copper in swine feeding. J. Anim. Sci. 19:249–259. Hoekstra, W. G. 1970. The complexity of dietary factors affecting zinc nutrition and metabolism in chicks and swine. Pp. 347–353 in Trace Element Metabolism in Animals, C. F. Mills, ed. Edinburgh: E. & S. Livingstone. Hoekstra, W. G., P. K. Lewis, P. H. Phillips, and R. H. Grummer. 1956. The relationship of parakeratosis, supplemental calcium and zinc to the zinc content of certain body components of swine. J. Anim. Sci. 15:752–764. Hoekstra, W. G., E. C. Faltin, C. W. Lin, H. F. Roberts, and R. H. Grummer. 1967. Zinc deficiency in reproducing gilts fed a diet high in calcium and its effect on tissue zinc and blood serum alkaline phosphatase. J. Anim. Sci. 26:1348–1357. Honeyfield, D. C., and J. A. Froseth. 1985. Effects of dietary sodium and chloride on growth, efficiency of feed utilization, plasma electrolytes and plasma basic amino acids in young pigs. J. Nutr. 115:1366–1371. Honeyfield, D. C., J. A. Froseth, and R. J. Barke. 1985. Dietary sodium and chloride levels for growing-finishing pigs. J. Anim. Sci. 60:691–698. Hsu, F. S., L. Krook, W. G. Pond, and J. R. Duncan. 1975. Interactions of dietary calcium with toxic levels of lead and zinc in pigs. J. Nutr. 105:112–118. Huang, K. C., and G. L. Allee. 1981. Bioavailability of phosphorus in selected feedstuffs for young chick and pigs. J. Anim. Sci. 53(Suppl. 1):248 (Abstr.). Huck, D. W., and A. J. Clawson. 1976. Cobalt toxicity in pigs. J. Anim. Sci. 43:253. (Abstr.) Hughes, E. H., and N. R. Ittner. 1942. The potassium requirement of growing pigs. J. Agric. Res. 64:189–192. Izquierdo, O. A., and D. H. Baker. 1986. Bioavailability of copper in pig feces. Can. J. Anim. Sci. 66:1145–1148. Jensen, A. H., S. W. Terrill, and D. E. Becker. 1961. Response of the young pig to levels of dietary potassium. J. Anim. Sci. 20:464–467.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Johnson, S. R. 1944. Studies with swine on low manganese rations of natural foodstuffs. J. Anim. Sci. 3:136–142. Jongbloed, A. W. 1987. Phosphorus in the Feeding of Pigs: Effect of Diet on the Absorption and Retention of Phosphorus by Growing Pigs. Instituut voor Veevoedingsanderzoek. Lelystad. XVI, 343 pp. Jongbloed, A. W., and P. A. Kemme. 1990. Effect of pelleting mixed feeds on phytase activity and the apparent absorbability of phosphorus and calcium in pigs. Anim. Feed Sci. Tech. 28:233–242. Jongbloed, A. W., Z. Mroz, and P. A. Kemme. 1992. The effect of supplementary Aspergillus niger phytase in diets for pigs on concentration and apparent digestibility of dry matter, total phosphorus, and phytic acid in different sections of the alimentary tract. J. Anim. Sci. 70:1159–1168. Jongbloed, A. W., Z. Mroz, P. A. Kemme, C. Geerse, and Y. Van Der Honing. 1993. The effect of dietary calcium levels on microbial phytase efficacy in growing pigs. J. Anim. Sci. 71(Suppl. 1):166 (Abstr.). Kadis, S., F. A. Udeze, J. Polanco, and D. W. Dreesen. 1984. Relationship of iron administration to susceptibility of newborn pigs to enterotoxic colibacillosis. Am. J. Vet. Res. 45:255–259. Kalinowski, J., and E. R. Chavez. 1984. Effect of low dietary zinc during late gestation and early lactation on the sow and neonatal piglets. Can. J. Anim. Sci. 64:749–758. Kalkus, J. W. 1920. A study of goiter and associated conditions in domestic animals. Wash. Agric. Exp. Stn. Bull. 156:1–48. Kavanagh, N. T. 1992. The effect of feed supplemented with zinc oxide on the performance of recently weaned pigs. P. 616 in 12th Proc. Internat. Pig Veterinary Soc. Congr. Kayongo-Male, H., D. E. Ullrey, and E. R. Miller. 1975. The Mn requirement of the baby pig from sows fed a low Mn diet. East Afr. Agric. For. J. 41(2):157–164. Kernkamp, H. C. H., and E. F. Ferrin. 1953. Parakeratosis in swine. J. Am. Vet. Med. Assoc. 123:217–220. Kernkamp, H. C. H., A. J. Clawson, and R. H. Ferneyhough. 1962. Preventing iron-deficiency anemia in baby pigs. J. Anim. Sci. 21:527–532. Kemme, P. A., A. W. Jongbloed, Z. Mroz, and M. Makinen. 1995. Apparent ileal digestibility of protein and amino acids from a maize-soybean meal diet with or without extrinsic phytate and phytase in pigs. In International Symposium on Nutrient Management of Food Animals to Enhance the Environment. Abstracts, P. 6, June 4–8. Kesel, G. A., J. W. Knight, E. T. Kornegay, J. P. Veit, and D. R. Notter. 1983. Restricted energy and elevated calcium and phosphorus intake for boars during growth. 1. Feedlot performance and bone characteristics. J. Anim. Sci. 57:82–98. Ketaren, P. P., E. S. Batterham, and D. J. Farrell. 1989. Dietary phosphorus levels and calcium available phosphorus ratio for growing pigs. Pp. 155–163 in Recent Advances in Animal Nutrition in Australia, D. J. Farrell, ed. Armidale, Australia, Dept. of Biochem., Microbiol. and Nutr., Univ. of New England. Ketaren, P. P., E. S. Batterham, E. B. Dettmann, and D. J. Farrell. 1993. Phosphorus studies in pigs. 3. Effect of phytase supplementation on the digestibility and availability of phosphorus in soy-bean meal for grower pigs. Br. J. Nutr. 70:289–311. Kirchgessner, M., and E. Grassman. 1970. The dynamics of copper absorption. Pp. 277–287 in Trace Element Metabolism in Animals, C. F. Mills, ed. Edinburgh: E. & S. Livingstone. Kirchgessner, M., H. Mader, and E. Grassman. 1980. Zur Fruchtbarkeitsleistung von Saven bei unterschiedlicher Cu-Versorgung. Zuchtungskunde 52:46–53. Kirchgessner, M., D. A. Roth-Maier, and R. Sporl. 1981. Untersuchungen zum Trachtigkeitsanabolismus der spurenelemente Kupfer, Zin, Nickel und Mangan bei Zuchtsaver. Arch. Tierernahr. 31:21–34. Klasing, K. C., C. D. Knight, and D. M. Forsyth. 1980. Effects of iron on the anti-coli capacity of sow’s milk in vitro and in ligated intestinal segments. J. Nutr. 110:1914–1921.

63

Kline, E. A., J. Kastelic, C. C. Ashton, P. G. Homeyer, L. Quinn, and D. V. Catron. 1954. The effect on the growth performance of young pigs of adding cobalt, vitamin B12 and antibiotics to semipurified rations. J. Nutr. 53:543–555. Klosterman, E. W., W. E. Dinusson, E. L. Lasley, and M. L. Buchanan. 1950. Effect of trace minerals on growth and fattening of swine. Science 112:168–169. Knight, D. D., K. C. Klasing, and D. M. Forsyth. 1983. E. coli growth in serum of iron dextran-supplemented pigs. J. Anim. Sci. 57:387–395. Koch, M. E., D. C. Mahan, and J. R. Corley. 1984. An evaluation of various biological characteristics in assessing low phosphorus intake in weanling swine. J. Anim. Sci. 59:1546–1556. Koprinski, J. S., J. Leibholz, and L. Bryden. 1989. Biotin studies in pigs. Biotin availability in feedstuffs for pigs and chickens. Br. J. Nutr. 62:773–780. Kornegay, E. T. 1972a. Availability of iron contained in defluorinated phosphate. J. Anim. Sci. 34:569–572. Kornegay, E. T. 1972b. Supplementation of lysine, ammonium polyphosphate and urea in diets for growing-finishing pigs. J. Anim. Sci. 34:55–63. Kornegay, E. T. 1985. Calcium and Phosphorus in Animal Nutrition. Pp. 1–106 in Calcium and Phosphorus in Animal Nutrition. West Des Moines, Iowa: National Feed Ingredients Association. Kornegay, E. T. 1996. Nutritional, Environmental and Economical Considerations for Using Phytase in Pig and Poultry Diets. Pp. 279–304 in International Symposium on Nutrient Management of Food Animals to Enhance the Environment, E. T. Kornegay, ed. Boca Raton, Fla.: CRC Press, Inc. Kornegay, E. T., and B. Kite. 1983. Phosphorus in swine. VI. Utilization of nitrogen, calcium and phosphorus and reproductive performance of gravid gilts fed two dietary phosphorus levels for five parities. J. Anim. Sci. 57:1463–1473. Kornegay, E. T., and H. Qian. 1994. Effectiveness of Natuphost phytase as influenced by dietary phosphorus for improving the bioavailability of phytate phosphorus in a corn-soybean meal based diet fed to young pigs. J. Anim. Sci. 72(Suppl. 1):330 (Abstr.). Kornegay, E. T. and J. S. Radcliffe. 1997. Relative bioavailability of phosphorus sources with different solubilities in neutral ammonium citrate (NAC) for young pigs. J. Anim. Sci. 75(Suppl. 1):188 (Abstr.). Kornegay, E. T., and H. R. Thomas. 1981. Phosphorus in swine. II. Influence of dietary calcium and phosphorus levels and growth rate on serum minerals, soundness scores and bone development in barrows, gilts and boars. J. Anim. Sci. 52:1049–1059. Kornegay, E. T., H. R. Thomas, and T. N. Meacham. 1973. Evaluation of dietary calcium and phosphorus for reproducing sows housed in total confinement on concrete or in dirt lots. J. Anim. Sci. 37:493–500. Kornegay, E. T., H. R. Thomas, and J. L. Baker. 1981a. Phosphorus in swine. IV. Influence of dietary calcium and phosphorus and protein levels on feedlot performance, serum minerals, bone development and soundness scores in boars. J. Anim. Sci. 52:1070–1084. Kornegay, E. T., H. R. Thomas, J. H. Carter, L. B. Allen, C. C. Brooks, and K. H. Hinklemann. 1981b. Phosphorus in swine. V. Interrelationships of various feedlot performance, serum minerals, structural soundness and bone parameters in barrows, boars and gilts. J. Anim. Sci. 52:1085–1090. Kornegay, E. T., H. P. Veit, J. W. Knight, D. R. Notter, H. S. Bartlett, and D. F. Calabotta. 1983. Restricted energy intake and elevated calcium and phosphorus intake for boars during growth. II. Foot and leg measurements and toe and soundness scores. J. Anim. Sci. 57:1182–1199. Kornegay, E. T., B. G. Diggs, O. M. Hale, D. L. Handlin, J. P. Hitchcock, and R. A. Barczwski. 1984. Reproductive performance of sows fed elevated calcium and phosphorus levels during growth and development. A cooperative study. Report S-145 of the Committee on Nutri-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

64

Nutrient Requirements of Swine

tional Systems for Swine to Increase Reproductive Efficiency. J. Anim. Sci. 59(Suppl. 1):253 (Abstr.). Kornegay, E. T., J. B. Meldrum, G. Schurig, M. D. Lindemann, and F. C. Gwazdauskas. 1986. Lack of influence of nursery temperature on the response of weanling pigs to supplemental vitamins C and E. J. Anim. Sci. 63:484–491. Kornegay, E. T., P. H. G. van Heugten, M. D. Lindemann and D. J. Blodgett. 1989. Effects of biotin and high copper levels on performance and immune response of weanling pigs. J. Anim. Sci. 67:1471–1477. Kornegay, E. T., M. D. Lindemann, and H. S. Bartlett. 1991. The influence of sodium supplementation of two phosphorus sources on performance and bone mineralization of growing-finishing swine evaluated at two geographical locations. Can. J. Anim. Sci. 71:537–547. Kornegay, E. T., J. L. Evans, and V. Ravidndran. 1994. Effects of diet acidity and protein level or source of calcium on the performance, gastrointestinal content measurements, bone measurements and carcass composition of gilt and barrow weanling pigs. J. Anim. Sci. 72:2670–2680. Kornegay, E. T., W. Zhou, J. W. G. M. Swinkels, and C. R. Risley. 1995. Characterization of cobalt–copper antagonism in the study of copperstimulated growth in weanling pigs. J. Anim. Feed Sci. 4:21–33. Krider, J. L., J. L. Albright, M. P. Plumlee, J. H. Conrad, C. L. Sinclair, L. Underwood, R. G. Jones, and R. B. Harrington. 1975. Magnesium supplementation, space and docking effects on swine performance and behavior. J. Anim. Sci. 40:1027–1033. Ku, P. K., D. E. Ullrey, and E. R. Miller. 1970. Zinc deficiency and tissue nucleic acid and protein concentration. Pp. 158–164 in Trace Element Metabolism in Animals, C. F. Mills, ed. Edinburgh: E. & S. Livingstone. Ku, P. K., W. T. Ely, A. W. Groce, and D. E. Ullrey. 1973. Natural dietary selenium, a-tocopherol and effect on tissue selenium. J. Anim. Sci. 37:501–505. Kuhajek, E. J., and G. F. Andelfinger. 1970. A new source of iodine for salt blocks. J. Anim. Sci. 31:51–58. Kulwich, R., S. L. Hansard, C. L. Comar, and G. K. Davis. 1953. Copper, molybdenum, and zinc interrelationships in rats and swine. Proc. Soc. Exp. Biol. Med. 84:487. Lannek, N., P. Lindberg, and G. Tollerz. 1962. Lowered resistance to iron in vitamin E-deficient piglets and mice. Nature 195:1006–1007. Larsen, H. J., and S. Tollersrud. 1981. Effect of dietary vitamin E and selenium on photohemagglutinin response of the pig lymphocytes. Res. Vet. Sci. 31:301–305. Leach, R. M., Jr., and A. M. Muenster. 1962. Studies on the role of manganese in bone formation. 1. Effect upon the mucopolysaccharide content of chick bone. J. Nutr. 78:51–56. Lei, X. G., P. K. Ku, E. R. Miller, and M. T. Yokoyama. 1993a. Supplementing corn–soybean meal diets with microbial phytase linearly improves phytate phosphorus utilization by weanling pigs. J. Anim. Sci. 71:3359–3367. Lei, X. G., P. K. Ku, E. R. Miller, D. E. Ullrey, and M. T. Yokoyama. 1993b. Supplemental microbial phytase improves bioavailability of dietary zinc to weanling pigs. J. Nutr. 123:1117–1123. Lei, X. G., P. K. Ku, E. R. Miller, M. T. Yokoyama, and D. E. Ullrey. 1994. Calcium level affects the efficacy of supplemental microbial phytase in corn–soybean meal diets of weanling pigs. J. Anim. Sci. 72:139–143. Leibholz, J. M., V. C. Speer, and V. W. Hays. 1962. Effect of dietary manganese on baby pig performance and tissue manganese levels. J. Anim. Sci. 21:772–776. Leibholz, J. M., J. T. McCall, V. W. Hays, and V. C. Speer. 1966. Potassium, protein and basic amino acid relationships in swine. J. Anim. Sci. 25:37–43. LeMieux, F. M., L. V. Ellison, T. L. Ward, L. L. Southern, and T. D. Bidner. 1995. Excess dietary zinc for pigs weaned at 28 days. J. Anim. Sci. 73(Suppl. 1):72 (Abstr.).

Lepine, A. J., E. T. Kornegay, D. R. Notter, H. P. Veit, and J. W. Knight. 1985. Foot and leg measurements, toe lesions, soundness scores and feedlot performance of crossbred boars as influenced by nutrition and age. Can. J. Anim. Sci. 65:459–472. Lewis, A. J., P. S. Miller, and C. K. Wolverton. 1995. Bioavailability of iron in iron methionine for weanling pigs. J. Anim. Sci. 73(Suppl. 1):172 (Abstr.). Lewis, P. K., Jr., W. C. Hoekstra, R. H. Grummer, and P. H. Phillips. 1956. The effects of certain nutritional factors including calcium, phosphorus and zinc on parakeratosis. J. Anim. Sci. 15:741–751. Lewis, P. K., Jr., R. H. Grummer, and W. G. Hoekstra. 1957a. The effect of method of feeding upon the susceptibility of the pig to parakeratosis. J. Anim. Sci. 16:927–936. Lewis, P. K., Jr., W. G. Hoekstra, and R. H. Grummer. 1957b. Restricted calcium feeding versus zinc supplementation for the control of parakeratosis in swine. J. Anim. Sci. 16:578–588. Libal, G. W., E. R. Peo, Jr., R. P. Andrews, and P. E. Vipperman, Jr. 1969. Levels of calcium and phosphorus for growing-finishing swine. J. Anim. Sci. 28:331–335. Lillie, R. J., and L. T. Frobish. 1978. Effect of copper and iron supplements on performance and hematology of confined sows and their progeny through four reproductive cycles. J. Anim. Sci. 46:678–685. Lindberg, P., and N. Lannek. 1965. Retention of selenium in kidneys, liver and striated muscle after prolonged feeding of therapeutic amounts of sodium selenite to pigs. Acta Vet. Scand. 6:217–223. Lindberg, P., and M. Siren. 1963. Selenium concentration in kidneys of normal pigs and pigs affected with nutritional muscular dystrophy and liver dystrophy (hepatosis dietetica). Life Sci. 2:326–330. Lindberg, P., and M. Siren. 1965. Fluorometric selenium determinations in the liver of normal pigs and in pigs affected with nutritional muscular dystrophy and liver dystrophy. Acta Vet. Scand. 6:59–64. Lindemann, M. D., A. F. Harper, and E. T. Kornegay. 1995a. Further assessment of the effects of supplementation of chromium from chromium picolinate on fecundity in swine. J. Anim. Sci. 73(Suppl. 1):185 (Abstr.). Lindemann, M. D., C. M. Wood, A. F. Harper, E. T. Kornegay, and R. A. Anderson. 1995b. Dietary chromium picolinate additions improve gain:feed and carcass characteristics in growing-finishing pigs and increase litter size in reproducing sows. J. Anim. Sci. 73:457–465. Linkenheimer, W. H., E. L. Patterson, R. A. Milstrey, J. A. Brockman, and D. E. Johnson. 1960. Preparation and biological testing of a parenteral iron preparation. J. Anim. Sci. 19:763–768. Liptrap, D. O., E. R. Miller, D. E. Ullrey, D. L. Whitenack, B. L. Schoepke, and R. W. Luecke. 1970. Sex influence on the zinc requirement of developing swine. J. Anim. Sci. 30:736–741. Lolas, G. M., N. Palamidis, and P. Markakis. 1976. The phytic acid-total phosphorus relationship in barley, oats, soybeans and wheat. Cereal Chem. 53:867–871. Lowry, K. R., D. C. Mahan, and J. R. Corley. 1985a. Effect of dietary calcium on selenium retention in postweaning swine. J. Anim. Sci. 60:1429–1437. Lowry, K. R., D. C. Mahan, and J. R. Corley. 1985b. Effect of dietary phosphorus on selenium retention in postweaning swine. J. Anim. Sci. 60:1438–1446. Luecke, R. W., J. A. Hoefer, W. G. Brammell, and F. Thorp, Jr. 1956. Mineral interrelationships in parakeratosis of swine. J. Anim. Sci. 15:247–251. Luecke, R. W., J. A. Hoefer, W. S. Brammell, and D. A. Schmidt. 1957. Calcium and zinc in parakeratosis of swine. J. Anim. Sci. 16:3–11. Luo, X. G., and C. R. Dove. 1996. Effect of dietary copper and fat on nutrient utilization, digestive enzyme activities, and tissue mineral levels in weanling pigs. J. Anim. Sci. 74:1888–1896. Madubuike, F. N., and R. E. Austic. 1989. Lysine, potassium and chloride in young growing pigs. J. Anim. Prod. Res. 9:1–10.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Madubuike, F. N., C. C. Calvert, and R. E. Austic. 1980. Lysine-cation interrelationship in the pig. J. Anim. Sci. 51(Suppl. 1):210 (Abstr.). Mahan, D. C. 1982. Dietary calcium and phosphorus levels for weanling swine. J. Anim. Sci. 54:559–564. Mahan, D. C. 1985. Effect of inorganic selenium supplementation on selenium retention in postweaning swine. J. Anim. Sci. 61:173–178. Mahan, D. C. 1995. Selenium metabolism in animals: What role does selenium yeast have? Pp. 257–267 in Biotechnology in the Feed Industry. Proc. Alltechs 11th Annual Symposium. T. P. Lyons and K. A. Jacques, eds. United Kingdom: Nottingham University Press. Mahan, D. C., and A. W. Fetter. 1982. Dietary calcium and phosphorus levels for reproducing sows. J. Anim. Sci. 54:285–291. Mahan, D. C., and Y. Y. Kim. 1996. Effect of inorganic selenium at two dietary levels on reproductive performance and tissue selenium concentrations in first-parity gilts and their progeny. J. Anim. Sci. 74:2711–2718. Mahan, D. C., and P. L. Magee. 1991. Efficacy of dietary sodium selenite and calcium selenite provided in the diet at approved, marginally toxic, and toxic levels to growing swine. J. Anim. Sci. 69:4722–4725. Mahan, D. C., and A. L. Moxon. 1978a. Effect of adding inorganic or organic selenium sources to the diets of young swine. J. Anim. Sci. 47:456–466. Mahan, D. C., and A. L. Moxon. 1978b. Effect of increasing the level of inorganic selenium supplementation in the postweaning diets of swine. J. Anim. Sci. 46:384–390. Mahan, D. C., and A. L. Moxon. 1984. Effect of inorganic selenium supplementation on selenosis in postweaning swine. J. Anim. Sci. 58:1216–1221. Mahan, D. C., and N. A. Parrett. 1996. Evaluating the efficacy of seleniumenriched yeast and sodium selenite on tissue selenium retention and serum glutathione peroxidase activity in grower and finisher diets. J. Anim. Sci. 74:2967–2974. Mahan, D. C., J. E. Jones, J. H. Cline, R. F. Cross, H. S. Teague, and A. P. Crifo, Jr. 1973. Efficacy of selenium and vitamin E injections in the prevention of white muscle disease in young swine. J. Anim. Sci. 36:1104–1108. Mahan, D. C., L. H. Penhale, J. H. Cline, A. L. Moxon, A. W. Fetter, and J. T. Yarrington. 1974. Efficacy of supplemental selenium in reproductive diets on sow and progeny performance. J. Anim. Sci. 39:536–543. Mahan, D. C., A. L. Moxon, and J. H. Cline. 1975. Efficacy of supplemental selenium in reproductive diets on sow and progeny serum and tissue selenium values. J. Anim. Sci. 40:624–631. Mahan, D. C., A. L. Moxon, and M. Hubbard. 1977. Efficacy of inorganic selenium supplementation to sow diets on resulting carry-over to their progency. J. Anim. Sci. 45:738–746. Mahan, D. C., K. E. Ekstrom, and A. W. Fetter. 1980. Effect of dietary protein, calcium and phosphorus for swine from 7 to 20 kilograms body weight. J. Anim. Sci. 50:309–314. Mahan, D. C., E. A. Newton, and K. R. Cera. 1996a. Effect of supplemental sodium chloride, sodium phosphate, or hydrochloric acid in starter pig diets containing dried whey. J. Anim. Sci. 74:1217–1222. Mahan, D. C., E. M. Weaver, and L. E. Russell. 1996b. Improved postweaning pig performance responses by adding NaCl or HCl to diets containing animal plasma. J. Anim. Sci. 74(Suppl. 1):58 (Abstr). Maner, J. H., W. G. Pond, and R. S. Lowrey. 1959. Effect of method and level of iron administration on growth, hemoglobin and hematocrit of suckling pigs. J. Anim. Sci. 18:1373–1377. Manners, M. J., and M. R. McCrea. 1964. Estimates of the mineral requirements of 2-day weaned piglets derived from data on mineral retention by sow-reared piglets. Ann. Zootechnol. 13:29–38. Matrone, G., R. H. Hartman, and A. J. Clawson. 1959. Studies of a manganese-iron antagonism in the nutrition of rabbits and baby pigs. J. Nutr. 67:309–317.

65

Matrone, G., E. L. Thomason, and C. R. Bunn. 1960. Requirement and utilization of iron by the baby pig. J. Nutr. 72:459–465. Maxson, P. F., and D. C. Mahan. 1983. Dietary calcium and phosphorus levels for growing swine from 18 to 57 kilograms body weight. J. Anim. Sci. 56:1124–1134. Maxson, P. F., and D. C. Mahan. 1986. Dietary calcium and phosphorus for lactating swine at high and average production levels. J. Anim. Sci. 63:1163–1172. Mayo, R. H., M. P. Plurnlee, and W. M. Beeson. 1959. Magnesium requirement of the pig. J. Anim. Sci. 18:264–273. McCance, R. A., and E. M. Widdowson. 1944. Activity of the phytase in different cereals and its resistance to dry heat. Nature 153:650. McCarrison, R. 1933. The goitrogenic action of soybean and groundnut. Indian J. Med. Res. 7:189. McCully, G. A., G. M. Hill, J. E. Link, R. L. Weavers, M. S. Carlson and D. W. Rozeboom. 1995. Evaluation of zinc sources for the newly weaned pig. J. Anim. Sci. 74(Suppl. 1):72 (Abstr.). McDonald, F. F., D. Dunlop, and C. M. Bates. 1955. An effective treatment for anemia of piglets. Br. Vet. J. 111:403–407. Menehan, L. A., P. A. Knapp, W. G. Pond, and J. R. Jones. 1963. Response of early-weaned pigs to variations in dietary calcium level with and without lactose. J. Anim. Sci. 22:501–505. Mertz, W. 1993. Chromium in human nutrition: A review. J. Nutr. 123:626–633. Meyer, J. H., R. H. Grummer, P. H. Phillips, and G. Bohstedt. 1950. Sodium, chlorine, and potassium requirements of growing pigs. J. Anim. Sci. 9:300–306. Meyer, W. R., D. C. Mahan, and A. L. Moxon. 1981. Value of dietary selenium and vitamin E for weanling swine as measured by performance and tissue selenium and glutathione peroxidase activities. J. Anim. Sci. 52:302–311. Michel, R. L., C. K. Whitehair, and K. K. Keahey. 1969. Dietary hepatic necrosis associated with selenium-vitamin E deficiency in swine. J. Am. Vet. Med. Assoc. 155:50–59. Mijada, V. S., and T. R. Cline. 1983. Lysine and potassium supplementation of low protein pig starter diets. J. Anim. Sci. 57(Suppl. 1):87 (Abstr.). Miller, E. R. 1975. Utilization of inorganic sulfate by growing-finishing swine. Mich. Agric. Exp. Stn. Res. Rep. 289:100–104. Miller, E. R. 1980. Bioavailability of minerals. P. 144 in Proc. Minnesota Nutrition Conference. St. Paul: University of Minnesota Press. Miller, E. R. 1991. Iron, copper, zinc, manganese, and iodine in swine nutrition. Pp. 267–284 in Swine Nutrition, E. R. Miller, D. E. Ullrey, and A. J. Lewis, eds. Stoneham, MA: Butterworth-Heinemann Publishing. Miller, E. R., and J. A. Froseth. 1982. Interaction of lysine and potassium in swine nutrition. P. 47 in Proc. Fifth Annual International Minerals Conference. Mundelein, IL: International Minerals and Chemical Corp. Miller, E. R., D. E. Ullrey, C. L. Zutaut, B. V. Baltzer, D. A. Schmidt, J. A. Hoefer, and R. W. Luecke. 1962. Calcium requirement of the baby pig. J. Nutr. 77:7–16. Miller, E. R., D. E. Ullrey, C. L. Zutaut, B. V. Baltzer, D. A. Schmidt, J. A. Hoefer, and R. W. Luecke. 1964a. Phosphorus requirement of the baby pig. J. Nutr. 82:34–39. Miller, E. R., D. E. Ullrey, C. L. Zutaut, J. A. Hoefer, and R. W. Luecke. 1964b. Mineral balance studies with the baby pig: Effects of dietary phosphorus level upon calcium and phosphorus balance. J. Nutr. 82:111–114. Miller, E. R., D. E. Ullrey, C. L. Zutaut, B. V. Baltzer, D. A. Schmidt, J. A. Hoefer, and R. W. Luecke. 1965a. Magnesium requirement of the baby pig. J. Nutr. 85:13–20. Miller, E. R., D. E. Ullrey, C. L. Zutaut, J. A. Hoefer, and R. W. Luecke. 1965b. Comparisons of casein and soy proteins upon mineral balance and vitamin D2 requirement of the baby pig. J. Nutr. 85:347–353.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

66

Nutrient Requirements of Swine

Miller, E. R., D. E. Ullrey, C. L. Zutaut, J. A. Hoefer, and R. W. Luecke. 1965c. Mineral balance studies with the baby pig: Effects of dietary magnesium level upon calcium, phosphorus, and magnesium balance. J. Nutr. 86:209–212. Miller, E. R., D. E. Ullrey, C. L. Zutaut, J. A. Hoefer, and R. W. Luecke. 1965d. Mineral balance studies with the baby pig: Effects of dietary vitamin D2 level upon calcium, phosphorus, and magnesium balance. J. Nutr. 85:255–258. Miller, E. R., R. W. Luecke, D. E. Ullrey, B. V. Baltzer, B. L. Bradley, and J. A. Hoefer. 1968. Biochemical, skeletal and allometric changes due to zinc deficiency in the baby pig. J. Nutr.95:278–286. Miller, E. R., D. O. Liptrap, and D. E. Ullrey. 1970. Sex influence on zinc requirement of swine. Pp. 377–379 in Trace Element Metabolism in Animals, C. F. Mills, ed. Edinburgh: E. & S. Livingstone. Miller, E. R., H. D. Stowe, P. K. Ku, and G. M. Hill. 1979. Copper and zinc in swine nutrition. P. 109 in National Feed Ingredients Association Literature Review on Copper and Zinc in Animal Nutrition. West Des Moines, Iowa: National Feed Ingredients Association. Miller, E. R., M. J. Parsons, D. E. Ullrey, and P. K. Ku. 1981a. Bioavailability of iron from ferric choline citrate and a ferric copper cobalt choline citrate complex for young pigs. J. Anim. Sci. 52:783–787. Miller, E. R., J. Skomial, P. K. Ku, and M. C. Hogberg. 1981b. Supplemental potassium in low lysine diets for swine. J. Anim. Sci. 53(Suppl. 1):90 (Abstr.). Miller, E. R., G. L. Waxler, P. K. Ku, D. E. Ullrey, and C. K. Whitehair. 1982. Iron requirements of baby pigs reared in germ-free or conventional environments on a condensed milk diet. J. Anim. Sci. 54:106–115. Miller, W. T. 1938. Toxicity of selenium fed to swine in the form of sodium selenite. J. Agric. Res. 56:831–842. Miller, W. T., and K. T. Williams. 1940. Minimum lethal dose of selenium, as sodium selenite, for horses, mules, cattle and swine. J. Agric. Res. 60:163–173. Miracle, G. L., G. L. Cromwell, T. S. Stahly, and D. D. Kratzer. 1977. Availability of phosphorus in corn, wheat, and soybean meal for pigs. J. Anim. Sci. 45(Suppl. 1):101 (Abstr.). Miser, R., and R. Blair. 1986. Reproductive performance of gilts and sows as affected by induced biotin deficiency and subsequent dietary biotin supplementation. J. Anim. Physiol. Anim. Nutr. 55:196–208. Mollgaard, H. 1946. On phytic acid, its importance in metabolism and its enzymic cleavage in bread supplemented with calcium. Biochem. J. 40:589–603. Monegue, H. J., T. S. Stahly, and G. L. Cromwell. 1980. Effects of dietary phosphorus levels on sow reproduction and lactation performance. J. Anim. Sci. 51(Suppl. 1):213 (Abstr.). Mongin, P. 1981. Recent advances in dietary cation-anion balance: applications in poultry. Proc. Nutr. Soc. 40:285–294. Mooney, K. W., and G. L. Cromwell. 1995. Effects of dietary chromium picolinate supplementation on growth, carcass characteristics, and accretion rates of carcass tissues in growing-finishing swine. J. Anim. Sci. 73:3351–3357. Mooney, K. W., and G. L. Cromwell. 1996. Effects of chromium picolinate on performance and tissue accretion in pigs with different lean gain potential. J. Anim. Sci. 74(Suppl. 1):65 (Abstr.) Mooney, K. W., and G. L. Cromwell. 1997. Efficacy of chromium picolinate and chromium chloride as potential carcass modifiers in swine. J. Anim. Sci. 75:2661–2671. Morgan, D. P., E. P. Young, I. P. Earle, R. J. Davey, and J. W. Stevenson. 1969. Effects of dietary calcium and zinc on calcium, phosphorus and zinc retention in swine. J. Anim. Sci. 29:900–905. Mraz, F. R., A. M. Johnson, and H. Patrick. 1958. Metabolism of cesium and potassium in swine as indicated by cesium-134 and potassium-42. J. Nutr. 64:541–548. Mroz, Z., A. W. Jongbloed, and P. A. Kemme. 1994. Apparent digestibility and retention of nutrients bound to phytate complexes as influenced

by microbial phytase and feeding regimen in pigs. J. Anim. Sci. 72:126–132. Mudd, A. J., W. C. Smith, and D. G. Armstrong. 1969. The influence of dietary concentration of calcium and phosphorus on their retention in the body of growing pigs. J. Agric. Sci. (Camb.) 73:189–195. Nasi, M. 1990. Microbial phytase supplementation for improving availability of plant phosphorus in the diet of the growing pig. J. Agric. Sci. Finl. 62:435–443. National Research Council. 1980. Mineral Tolerance of Domestic Animals. Washington, DC: National Academy Press. 577 pp. National Research Council. 1995. Nutrient Requirements of Laboratory Animals. 4th Ed. Washington, DC: National Academy Press. 173 pp. National Research Council. 1997. The Role of Chromium in Animal Nutrition. Washington, DC: National Academy Press. 80 pp. Nelson, T. S., L. W. Ferrara, and N. L. Storer. 1968. Phytate phosphorus content of feed ingredients derived from plants. Poult. Sci. 47:1372–1374. Nesheim, M. C., R. M. Leach, Jr., T. R. Zeigler, and J. A. Serafin. 1964. Interrelationships between dietary levels of sodium, chlorine and potassium. J. Nutr. 84:361–366. Newland, H. W., and G. K. Davis. 1961. Placental transfer of manganese in swine. J. Anim. Sci. 20:15–17. Newland, H. W., D. E. Ullrey, J. A. Hoefer, and R. W. Luecke. 1958. The relationship of dietary calcium to zinc metabolism in pigs. J. Anim. Sci. 17:886–892. Newton, G. L., and A. J. Clawson. 1974. Iodine toxicity: Physiological effects of elevated dietary iodine on pigs. J. Anim. Sci. 39:879–884. Nielsen, F. H. 1984. Ultratrace elements in nutrition. Annu. Rev. Nutr. 4:21–41. Nielsen, F. H. 1994. Chromium. Pp. 264–268 in Modern Nutrition in Health and Disease. 8th ed. M. E. Shils, J. A. Olson, and M. Shike, eds. Philadelphia: Lea & Febiger. Nielsen, H. E., V. Danielsen, M. G. Simesen, G. Grissel-Nielsen, W. Hjarde, T. Leth, and A. Basse. 1979. Selenium and vitamin E deficiency in pigs. Acta Vet. Scand. 20:276–288. Nilsson, P. O. 1960. Acute iron poisoning with myocardial degeneration in piglets. Nord. Veterinaermed. 12:113–119. Nimmo, R. D., E. R. Peo, Jr., J. D. Crenshaw, B. D. Moser, and A. J. Lewis. 1981a. Effects of level of dietary calcium-phosphorus during growth and gestation on calcium-phosphorus balance and reproductive performance of first-litter sows. J. Anim. Sci. 52:1343–1349. Nimmo, R. D., E. R. Peo, Jr., B. D. Moser, and A. J. Lewis. 1981b. Effect of level of dietary calcium-phosphorus during growth and gestation on performance, blood, and bone parameters of swine. J. Anim. Sci. 52:1330–1342. Norrdin, R. W., L. Krook, W. G. Pond, and E. F. Walker. 1973. Experimental zinc deficiency in weanling pigs on high- and low-calcium diets. Cornell Vet. 63:264–290. Nunes, C. S. 1993. Evaluation of phytase resistance in swine diets to different pelleting temperatures. Pp. 269–271 in Enzymes in Anim. Nutr. Proc. First Symp.-Kartause Ittingen, Switzerland. Nuoranne, P. J., R. P. Raunio, P. Saukko, and H. Karppanen. 1980. Metabolic effects of a low-magnesium diet in pigs. Br. J. Nutr. 44:53–60. Oberleas, D. 1983. The role of phytate in zinc bioavailability and homeostasis. Pp. 145–158 in Nutritional Bioavailability of Zinc, G. E. Inglett, ed. American Chemical Society Symposium Series No. 210. Washington, DC: American Chemical Society. Oberleas, D., and B. F. Harland. 1996. Impact of phytate on nutrient availability. Pp. 77–84 in Phytase in Animal Nutrition and Waste Management, M. B. Coelho, and E. T. Kornegay, eds. BASF Corporation, Mount Olive, NJ. Oberleas, D. M. E. Muhrer, and B. L. O’Dell. 1962. Effects of phytic acid on zinc availability and parakeratosis in swine. J. Anim. Sci. 21:57–61.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals O’Connor, D. L., M. F. Picciano, M. A. Roos, and R. A. Easter. 1989. Iron and folate utilization in reproducing swine and their progeny. J. Nutr. 119:1984–1991. O’Donovan, P. B., R. A. Pickett, M. P. Plumlee, and M. W. Beeson. 1963. Iron toxicity in the young pig. J. Anim. Sci. 22:1075–1080. O’Hara, P. J., A. P. Newman, and R. Jackson. 1960. Parakeratosis and copper poisoning in pigs fed a copper supplement. Aust. Vet. J. 36:225–229. Okonkwo, A. C., P. K. Ku, E. R. Miller, K. K. Keahey, and D. E. Ullrey. 1979. Copper requirement of baby pigs fed purified diets. J. Nutr. 109:939–948. Orstadius, K. 1960. Toxicity at a single subcutaneous dose of sodium selenite in pigs. Nature l88:1117. Orstadius, K., B. Wretlind, P. Lindberg. C. Nordstrom, and N. Lannek. 1959. Plasma transaminase and transferase activities in pigs affected with muscular and liver dystrophy. Zentralbl. Veterinaermed. 6:971–980. Osborne, J. C., and J. W. Davis. 1968. Increased susceptibility to bacterial endotoxin of pigs with iron-deficiency anemia. J. Am. Vet. Med. Assoc. 152:1630–1632. Owen, A. A., E. R. Peo, Jr., P. J. Cunningham, and B. D. Moser. 1973. Effect of EDTA on utilization of dietary zinc by G-F swine. J. Anim. Sci. 37:470–478. Page, T. G., L. L. Southern, T. L. Ward, and D. L. Thompson, Jr. 1993. Effect of chromium picolinate on growth and serum and carcass traits of growing-finishing pigs. J. Anim. Sci. 71:656–662. Pallauf, V. J., D. Hohler, and G. Rimbach. 1992a. Effect of microbial phytase supplementation to a maize-soya-diet on the apparent absorption on Mg, Fe, Cu, Mn and Zn and parameters of Zn-status in piglets. J. Anim. Physiol. Anim. Nutr. 68:1–9. Pallauf, V. J., D. Hohler, G. Rimbach, and H. Neusser. 1992b. Effect of microbial phytase supplementation to a maize-soy-diet on the apparent absorption of phosphorus and calcium in piglets. J. Anim. Physiol. Anim. Nutr. 67:30–40. Pallauf, J., G. Rimbach, S. Pippig, B. Schindler, and E. Most. 1994a. Effect of phytase supplementation to a phytate-rich diet based on wheat, barley and soya on the bioavailability of dietary phosphorus, calcium, magnesium, zinc and protein in piglets. Agribiol. Res. 47:39–48. Pallauf, J., G. Rimbach, S. Pippig, B. Schindler, and E. Most. 1994b. Dietary effect of phytogenic phytase and an addition of microbial phytase to a diet based on field beans, wheat, peas and barley on the utilization of phosphorus, calcium, magnesium, zinc and protein in piglets. Z. Ernahrungswiss 33:128–135. Parsons, M. J. 1979. The iron requirement of the neonatal pig from parental iron dextran. Mich. Agric. Exp. Stn. Res. Rep. 386:14. Partridge, I. G. 1981. A comparison of defluorinated rock phosphate and dicalcium phosphate, in diets containing either skim milk powder or soya bean meal as the main protein supplement, for early-weaned pigs. Anim. Prod. 32:67–73. Patience, J. F., and M. S. Wolynetz. 1990. Influence of dietary undetermined anion on acid-base status and performance in pigs. J. Nutr. 120:579–587. Patience, J. F., R. E. Austic, and R. D. Boyd. 1987. Effect of dietary electrolyte balance on growth and acid-base status in swine. J. Anim. Sci. 64:457–466. Patterson, D. S. P., W. M. Allen, D. C. Thurley, and J. T. Done. 1967. The role of tissue peroxidation in iron-induced myodegeneration of piglets. Biochem. J. 104:2P–3P. Patterson, D. S. P., W. M. Allen, S. Berrett, D. Sweasy, D. C. Thurley, and J. T. Done. 1969. A biochemical study of the pathogenesis of ironinduced myodegeneration of piglets. Zentralbl. Veterinaermed. 16:199–214.

67

Peeler, H. T. 1972. Biological availability of nutrients in feeds: Availability of major mineral ions. J. Anim. Sci. 35:695–712. Peo, E. R., Jr. 1976. Calcium in Swine Nutrition. West Des Moines, IA: National Feed Ingredient Association. 65 pp. Peo, E. R., Jr. 1991. Calcium, phosphorus, and vitamin D in swine nutrition. Pp. 165–182 in Swine Nutrition, E. R. Miller, D. E. Ullrey, and A. J. Lewis, eds. Stoneham, ME: Butterworth-Heinemann Publishing. Peo, E. R., Jr., G. W. Libal, G. F. Wehrbein, P. J. Cunningham, and P. E. Vipperman, Jr. 1969. Effect of dietary increments of calcium and phosphorus on G-F swine. J. Anim. Sci. 29:141(Abstr.). Peo, E. R., Jr., A. J. Lewis, and J. D. Crenshaw. 1982a. Effect of aluminum: fluorine ratios on the bioavailability of phosphorus from raw rock phosphate for G-F swine. J. Anim. Sci. 55(Suppl. 1):290 (Abstr.). Peo, E. R., Jr., A. J. Lewis, J. D. Crenshaw, and L. V. Himmelberg. 1982b. Effect of aluminum on the bioavailability of phosphorus from raw rock phosphate for G-F swine. J. Anim. Sci. 55(Suppl. 1):290 (Abstr.). Peplowski, M. A., D. C. Mahan, F. A. Murray, A. L. Moxon, A. H. Cantor, and K. E. Ekstrom. 1980. Effect of dietary and injectable vitamin E and selenium in weanling swine antigenically changed with sheep red blood cell. J. Anim. Sci. 51:344–351. Piatkowski, T. L., D. C. Mahan, A. H. Cantor, A. L. Moxon, J. H. Cline, and A. P. Grifo, Jr. 1979. Selenium and vitamin E in semipurified diets for gravid and nongravid gilts J. Anim. Sci. 48:1357–1365. Pickett, R. A., M. P. Plumlee, W. H. Smith, and W. M. Beeson. 1960. Oral iron requirement of the early-weaned pig. J. Anim. Sci. 19:1284. (Abstr.) Piper, R. C., J. A. Froseth, C. R. McDowell, G. H. Kroening, and I. A. Dyer. 1975. Selenium-vitamin E deficiency in swine fed peas (Pisum sativum). Am. J. Vet. Res. 36:273–281. Plumlee, M. P., D. M. Thrasher, W. M. Beeson, F. N. Andrews, and H. E. Parker. 1956. The effects of a manganese deficiency upon the growth, development and reproduction of swine. J. Anim. Sci. 15:352–368. Plumlee, M. P., C. E. Jordan, M. H. Kennington, and W. M. Beeson. 1958. Availability of the phosphorus from various phosphate materials for swine. J. Anim. Sci. 17:73–88. Pointillart, A., and L. Gueguen. 1978. Osteochondrose et faiblesse des pattes chez le porc. Ann. Biol. Anim. Bioch. Biophys. 18:201–210. Pointillart, A., N. Fontaine, and M. Thomasset. 1984. Phytate phosphorus utilization and intestinal phosphatases in pigs fed low phosphorus: Wheat or corn diets. Nutr. Rep. Int. 29:473–483. Pollmann, D. S., J. E. Smith, J. S. Stevenson, D. A. Schoneweis, and R. H. Hines. 1983. Comparison of gleptoferron with iron dextran for anemia prevention in young pigs. J. Anim. Sci. 56:640–644. Pond, W. G., and J. R. Jones. 1964. Effect of level of zinc in high-calcium diets on pigs from weaning through one reproductive cycle and on subsequent growth of their offspring. J. Anim. Sci. 23:1057–1060. Pond, W. G., R. S. Lowrey, J. H. Maner, and J. K. Loosli. 1961. Parenteral iron administration to sows during gestation or lactation. J. Anim. Sci. 20:747–750. Pond, W. G., J. R. Jones, and G. H. Kroening. 1964. Effect of level of dietary zinc and source and level of corn on performance and incidence of parakeratosis in weanling pigs. J. Anim. Sci. 23:16–20. Pond, W. G., P. Chapman, and E. Walker. 1966. Influence of dietary zinc, corn oil and cadmium on certain blood components, weight gain and parakeratosis in young pigs. J. Anim. Sci. 25:122–127. Pond, W. G., E. F. Walker, Jr., and D. Kirkland. 1975. Weight gain, feed utilization and bone and liver mineral composition of pigs fed high or normal Ca-P diets from weaning to slaughter weight. J. Anim. Sci. 41:1053–1056. Pond, W. G., E. F. Walker, Jr., and D. Kirkland. 1978. Effect of dietary Ca and P levels from 40 to 100 kg body weight on weight gain and bone and soft tissue mineral concentrations. J. Anim. Sci. 46:686–691.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

68

Nutrient Requirements of Swine

Pond, W. G., J. T. Yen, D. A. Hill, and W. E. Wheeler. 1981. Dietary Ca source and level: Effects on weanling pigs. J. Anim. Sci. 53(Suppl. 1):91 (Abstr.). Poulsen, H. D. 1989. Zinc oxide for pigs during weaning. Meddelelse No. 746. Statens Husdrybrugsforsoeq (Denmark). Prasad, A. S., D. Oberleas, P. Wolf, J. P. Horwitz, E. R. Miller, and R. W. Luecke. 1969. Changes in trace elements and enzyme activities in tissues of zinc-deficient pigs. Am. J. Clin. Nutr. 22:628–637. Prasad, A. S., D. Oberleas, E. R. Miller, and R. W. Luecke. 1971. Biochemical effects of zinc deficiency: Changes in activities of zinc-dependent enzymes and ribonucleic acid and deoxyribonucleic acid content of tissues. J. Lab. Clin. Med. 77:144–152. Prince, T. J., V. W. Hays, and G. L. Cromwell. 1975. Environmental effects of high copper pig manure on pasture for sheep. J. Anim. Sci. 41:326 (Abstr.). Prince, T. J., V. W. Hays, and G. L. Cromwell. 1984. Interactive effects of dietary calcium, phosphorus and copper on performance and liver copper stores of pigs. J. Anim. Sci. 58:356–361. Qian, H., E. T. Kornegay, and D. E. Conner, Jr. 1996. Adverse effects of wide calcium:phosphorus ratios on supplemental phytase efficacy for weanling pigs fed two dietary phosphorus levels. J. Anim. Sci. 74:1288–1297. Radcliffe, J. S., E. T. Kornegay, and D. E. Conner, Jr. 1995. The effect of phytase on calcium release in weanling pigs fed corn–soybean meal diets. J. Anim. Sci. 73(Suppl. 1):173 (Abstr.). Ramisz, A., A. Balicka-Laurans, and G. Ramisz. 1993. The influence of selenium on production, reproduction and health in pigs. Advances Ag. Sci. 2:67. Reinhart, G. A., and D. C. Mahan. 1986. Effect of various calcium: phosphorus ratios at low and high dietary phosphorus for starter, grower and finisher swine. J. Anim. Sci. 63:457–466. Rheaume, J. A., and E. R. Chavaz. 1989. Trace mineral metabolism in nongravid, gestating and lactating gilts fed two dietary levels of manganese. J. Trace Elem. Electrolytes Health Dis. 3:231–242. Rickes, E. L., N. G. Brink, F. R. Koniuszy, T. R. Wood, and K. Folkers. 1948. Vitamin B12, cobalt complex. Science 108:134. Ritchie, H. D., R. W. Luecke, B. V. Baltzer, E. R. Miller, D. E. Ullrey, and J. A. Hoefer. 1963. Copper and zinc interrelationships in the pig. J. Nutr. 79:117–123. Robinson, W. L. 1950. Cobalt with a soybean oil meal ration for pigs. J. Anim. Sci. 9:665. (Abstr.) Roof, M. D., and D. C. Mahan. 1982. Effect of carbadox and various dietary copper levels for weanling swine. J. Anim. Sci. 55:1109–1117. Roos, M. A., and R. A. Easter. 1986. Effect on sow and piglet performance of feeding a diet containing 250 ppm copper during lactation. J. Anim. Sci. 63(Suppl. 1):115 (Abstr.). Ross, R. D. 1983. The Biological Availability of Calcium and Phosphorus in Selected Feedstuffs for the Pig. Ph.D. Dissertation. University of Kentucky, Lexington. Ross, R. D., C. L. Cromwell, and T. S. Stahly. 1982. Biological availability of the phosphorus in regular and dehulled soybean meal for growing pigs. J. Anim. Sci. 55(Suppl. 1):93 (Abstr.). Ross, R. D., C. L. Cromwell, and T. S. Stahly. 1983. Biological availability of the phosphorus in high-moisture and pelleted corn. J. Anim. Sci. 57(Suppl. 1):96 (Abstr.). Ross, R. D., G. L. Cromwell, and T. S. Stahly. 1984. Effects of source and particle size on the biological availability of calcium in calcium supplements for growing pigs. J. Anim. Sci. 59:125–134. Rotruck, J. T., A. L. Pope, H. E. Canther, A. B. Swanson, D. C. Hafeman, and W. G. Hoekstra. 1973. Selenium: Biochemical role as a component of glutathione peroxidase. Science 179:588–590. Ruth, C. R., and J. F. Van Vleet. 1974. Experimentally induced seleniumvitamin E deficiency in growing swine: Selective destruction of type I skeletal muscle fibers. Am. J. Vet. Res. 35:237–244.

Rutledge, E. A., L. E. Hanson, and R. J. Meade. 1961. A study of the calcium requirements of pigs weaned at three weeks of age. J. Anim. Sci. 20:243–245. Rydberg, M. E., H. L. Self, T. Kowalczyk, and R. H. Grummer. 1959. The effect of prepartum intramuscular iron treatment of dams on litter hemoglobin levels. J. Anim. Sci. 18:415–419. Sansom, B. F., and P. T. Cleed. 1981. The ingestion of sow’s faeces by suckling piglets. Br. J. Nutr. 46:451–456. Sazzad, H. M., R. A. Vale, and E. P. Filgueiras. 1993. Bioavailability of copper from inorganic sources for weaning pigs. VII World Conf. on Anim. Prod., Edmonton, Canada, 94 (Abstr. #50). Schell, T. C., and E. T. Kornegay. 1996. Zinc concentration in tissues and performance of weanling pigs fed pharmacological levels of zinc from ZnO, Zn-methionine, Zn-lysine, and ZnSO 4 . J. Anim. Sci. 74:1584–1593. Shanklin, S. H., E. R. Miller, D. E. Ullrey, J. A. Hoefer, and R. W. Luecke. 1968. Zinc requirement of baby pigs on casein diets. J. Nutr. 96:101–108. Sharp, B. A., L. C. Young, and A. A. van Dreummel. 1972a. Dietary induction of mulberry heart disease and hepatosis dietetica in pigs. 1. Nutritional aspects. Can. J. Comp. Med. 36:371–376. Sharp, B. A., L. C. Young, and A. A. van Dreummel. 1972b. Effect of supplemental vitamin E and selenium in high–moisture corn diets on the incidence of mulberry heart disease and hepatosis dietetica in pigs. Can. J. Comp. Med. 36:393–397. Sheffy, B. E., and R. D. Schultz. 1979. Influence of vitamin E and selenium on immune response mechanisms. Fed. Proc. 38:2139–2143. Sihombing, D. T. H., G. L. Cromwell, and V. W. Hays. 1974. Effects of protein source, goitrogens and iodine level on performance and thyroid status of pigs. J. Anim. Sci. 39:1106–1112. Simesen, M. C., P. T. Jensen, A. Basse, C. Cissel–Nielsen, T. Leth, V. Danielsen, and H. E. Nielsen. 1982. Clinicopathologic findings in young pigs fed different levels of selenium, vitamin E and antioxidants. Acta Vet. Scand. 23:295–308. Simons, P. C. M., H. A. J. Versteegh, A. W. Jongbloed, P. A. Kemme, P. Slump, K. D. Bos, M. G. E. Wolters, R. F. Beudeker, and G. J. Verschoor. 1990. Improvement of phosphorus availability by microbial phytase in broilers and pigs. Br. J. Nutr. 64:525–540. Slatter, E. E. 1955. Mild iodine deficiency and losses of newborn pigs. J. Am. Vet. Med. Assoc. 127:149–152. Smith, II, J. W., M. D. Tokach, R. D. Goodband, J. L. Nelssen, W. B. Nessmith, Jr., K. Q. Owen, and B. T. Richert. 1995a. The effect of increasing zinc oxide supplementation on starter pig growth performance. J. Anim. Sci. 73(Suppl. 1):72. (Abstr). Smith, II, J. W., M. D. Tokach, R. D. Goodband, J. L. Nelssen, B. T. Richert, and S. S. Dritz. 1995b. Interaction between zinc oxide and copper sulfate on starter pig performance. J. Anim. Sci. 73(Suppl. 1):73. (Abstr). Smith, K. 1966. Inactivation of gossypol with mineral salts. Memphis, TN: National Cottonseed Production Association. Smith, W. H., M. P. Plumlee, and W. M. Beeson. 1958. Zinc requirement for growing swine. Science 128:1280–1281. Smith, W. H., M. P. Plumlee, and W. M. Beeson. 1962. Effect of source of protein on zinc requirement of the growing pig. J. Anim. Sci. 21:399–405. Soares, J. H. 1995. Calcium bioavailability. Pp. 95–118 in Bioavailability of Nutrients for Animals, C. B. Ammerman, D. H. Baker, and A. J. Lewis, eds. New York: Academic Press. Southern, L. L., and D. H. Baker. 1981. Effect of methionine or cysteine on cobalt toxicity in the chick. Poultry Sci. 60:1303–1308. Spallholz, J. E. 1980. Selenium: What role in immunity and immune cytoxicity. Pp. 103–117 in Proc. Second International Symposium on Selenium in Biology and Medicine, J. E. Spallholz, J. L. Martin, and H. E. Ganther, eds. Westport, CT: AVI Publishing Co.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minerals Spruill, D. C., V. W. Hays, and G. L. Cromwell. 1971. Effects of dietary protein and iron on reproduction and iron-related blood constituents in swine. J. Anim. Sci. 33:376–384. Stahly, T. S., C. L. Cromwell, and H. J. Monegue. 1980. Effects of the dietary inclusion of copper and (or) antibiotics on the performance of weanling pigs. J. Anim. Sci. 51:1347–1351. Stansbury, W. F., L. F. Tribble, and D. E. Orr, Jr. 1990. Effect of chelated copper sources on performance of nursery and growing pigs. J. Anim. Sci. 68:1318–1322. Stant, E. C., T. C. Martin, and W. V. Kassler. 1969. Potassium content of the porcine body and carcass at 23, 46, 68 and 91 kilograms live weight. J. Anim. Sci. 29:547–556. Stearns, D. M., J. J. Belbruno, and K. E. Wetterhahn. 1995. Chromium (III) picolinate produces chromosome damage in Chinese hamster ovary cells. FASEB J. 9:1643–1648. Steele, N. C., T. G. Althen, and L. T. Frobish. 1977. Biological activity of glucose tolerance factor in swine. J. Anim. Sci. 45:1341–1345. Stevenson, J W., and I. P. Earle. 1956. Studies on parakeratosis in swine. J. Anim. Sci. 15:1036–1045. Stober, C. R., C. L. Cromwell, and T. S. Stahly. 1979. Availability of phosphorus in corn and barley for the pig. J. Anim. Sci. 49(Suppl. 1):97 (Abstr.). Stober, C. R., G. L. Cromwell, and T. S. Stahly. 1980a. Biological availability of the phosphorus in cottonseed meal for growing pigs. J. Anim. Sci. 51(Suppl. 1):49 (Abstr.). Stober, C. R., G. L. Cromwell, and T. S. Stahly. 1980b. Biological availability of the phosphorus in oats, wheat middlings, and wheat bran for pigs. J. Anim. Sci. 51(Suppl. 1):80 (Abstr.). Stockland, W. L., and L. C. Blaylock. 1973. Influence of dietary calcium and phosphorus levels on the performance and bone characteristics of growing-finishing swine. J. Anim. Sci. 37:906–912. Stowe, H. D., and T. H. Herdt. 1992. Clinical assessment of selenium status of livestock. J. Anim. Sci. 70:3928–3933. Suomi, K., and T. Alaviuhkola. 1992. Responses to organic and inorganic selenium in the performance and blood selenium content of growing pigs. Ag. Sci. Finland 1:211. Suttle, N. F., and C. F. Mills. 1966a. Studies of toxicity of copper to pigs. I. Effects of oral supplements of zinc and iron salts on the development of copper toxicosis. Br. J. Nutr. 20:135–148. Suttle, N. F., and C. F. Mills. 1966b. Studies of toxicity of copper to pigs. II. Effect of protein source and other dietary components on the response to high and moderate intakes of copper. Br. J. Nutr. 20:149–161. Svajgr, A. J., E. R. Peo, Jr., and P. E. Vipperman, Jr. 1969. Effects of dietary levels of manganese and magnesium on performance of growing-finishing swine raised in confinement and on pasture. J. Anim. Sci. 29:439–443. Swinkels, J. W. G. M., E. T. Kornegay, W. Zhou, M. D. Lindemann, K. E. Webb, Jr., and M. W. A. Verstegen. 1996. Effectiveness of a zinc amino acid chelate and zinc sulfate in restoring serum and soft tissue zinc concentrations when fed to zinc-depleted pigs. J. Anim. Sci. 74:2420–2430. Taylor, T. G. 1965. The availability of the calcium and phosphorus of plant materials for animals. Proc. Nutr. Soc. 24:105–112. Teague, H. S., and L. E. Carpenter. 1951. The demonstration of copper deficiency in young growing pigs. J. Nutr. 43:389–399. Thacker, P. A. 1991. Effect of high levels of copper or dichlorvos during late gestation and lactation on sow productivity. Can. J. Anim. Sci. 71:227–248. Theuer, R. C., and W. C. Hoekstra. 1966. Oxidation of 14C-labeled carbohydrate, fat and amino acid substrates by zinc-deficient rats. J. Nutr. 89:448–454. Thomas, H. R., and E. T. Kornegay. 1981. Phosphorus in swine. I. Influence of dietary calcium and phosphorus levels and growth rate on

69

feedlot performance of barrows, gilts and boars. J. Anim. Sci. 52:1041–1048. Thompson, R. H., C. H. McMurray, and W. J. Blanchflower. 1976. The levels of selenium and glutathione peroxidase activity in blood of sheep, cows and pigs. Res. Vet. Sci. 20:229–231. Thoren-Tolling, K. 1975. Studies on the absorption of iron after oral administration in piglets. Acta Vet. Scand. Suppl. 54:1–121. Tokach, L. M., M. D. Tokach, R. D. Goodband, J. L. Nelssen, S. C. Henry, and T. A. Marsteller. 1992. Influence of zinc oxide in starter diets on pig performance. P. 411 in Proc. American Association of Swine Practitioners. Tonroy, B., M. P. Plumlee, J. H. Conrad, and T. R. Cline. 1973. Apparent digestibility of the phosphorus in sorghum grain and soybean meal for growing swine. J. Anim. Sci. 36:669–673. Trapp, A. L., K. K. Keahey, D. L. Whitenack, and C. K. Whitehair. 1970. Vitamin E-selenium deficiency in swine. Differential diagnosis and nature of field problem. J. Am. Vet. Med. Assoc. 157:289–300. Traylor, S. L., and G. L. Cromwell. 1998. Bioavailability of phosphorus in meat and bone meal for growing pigs. J. Anim. Sci. 76(Suppl. 2) Abstract no. 119, Midwestern Section meeting of the American Society of Animal Science, Des Moines, IA. Trotter, M., and C. L. Allee. 1979a. Availability of phosphorus in corn, soybean meal and wheat. J. Anim. Sci. 49(Suppl. 1):255 (Abstr.). Trotter, M., and G. L. Allee. 1979b. Availability of phosphorus in dry and high-moisture grain for pigs and chicks. J. Anim. Sci. 49(Suppl. 1):98 (Abstr.). Trotter, M., and G. L. Allee. 1979c. Effects of steam pelleting and extruding sorghum grain-soybean meal diets on phosphorus availability for swine. J. Anim. Sci. 49(Suppl. 1):255 (Abstr.). Tucker, H. F., and W. D. Salmon. 1955. Parakeratosis or zinc deficiency disease in the pig. Proc. Soc. Exp. Biol. Med. 88:613–616. Tunmire, D. L., D. E. Orr, Jr., and L. F. Tribble. 1983. Ammonium polyphosphate vs. dicalcium phosphate as a phosphorus supplement for growing-finishing swine. J. Anim. Sci. 57:632–637. Ullrey, D. E. 1992. Basis for regulation of selenium supplements in animal diets. J. Anim. Sci. 70:3922–3927. Ullrey, D. E. 1974. The selenium-deficiency problem in animal agriculture. Pp. 275–293 in Trace Element Metabolism in Animals, Volume 2, W. C. Hoekstra, J. W. Suttie, H. E. Ganther, and W. Mertz, eds. Baltimore: University Park Press. Ullrey, D. E., E. R. Miller, D. R. West, D. A. Schmidt, R. W. Seerley, J. A. Hoefer, and R. W. Luecke. 1959. Oral and parenteral administration of iron in the prevention and treatment of baby pig anemia. J. Anim. Sci. 18:256–263. Ullrey, D. E., E. R. Miller, O. A. Thompson, l. M. Ackermann, D. A. Schmidt, J. A. Hoefer, and R. W. Luecke. 1960. The requirement of the baby pig for orally administered iron. J. Nutr. 70:187–192. Ullrey, D. E., E. R. Miller, J. P. Hitchcock, P. K. Ku, R. L. Covert, J. Hegenauer, and P. Saltman. 1973. Oral ferric citrate vs. ferrous sulfate for prevention of baby pig anemia. Mich. Agric. Exp. Stn. Res. Rep. 232:34–38. Underwood, E. J. 1971. Trace Elements in Human and Animal Nutrition, Third edition. New York: Academic Press. Underwood, E. J. 1977. Trace Elements in Human and Animal Nutrition, Fourth edition. New York: Academic Press. van Kempen, G. J. M., P. van der Kerk, and A. H. M. Crimbergen. 1976. The influence of the phosphorus and calcium content of feeds on growth, feed conversion and slaughter quality and on the chemical, mechanical and histological parameters on the bone tissue of pigs. Neth. J. Agric. Sci. 24:120–139. van Vleet, J. F., K. B. Meyer, and H. J. Olander. 1973. Control of seleniumvitamin E deficiency in growing swine by parenteral administration of selenium-vitamin E preparations to baby pigs or to pregnant sows and their baby pigs. J. Am. Vet. Med. Assoc. 163:452–456.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

70

Nutrient Requirements of Swine

van Vleet, J. F., A. H. Rebar, and V. J. Ferns. 1977. Acute cobalt and isoproterenol cardiotoxicity in swine: Protection by selenium-vitamin E supplementation and potentiation by stress-susceptible phenotype. Am. J. Vet. Res. 38:991–1002. Venn, J. A. J., R. A. McCance, and E. M. Widdowson. 1947. Iron metabolism in piglet anemia. J. Comp. Pathol. Ther. 57:314–325. Veum, T. L., J. T. Gallo, W. G. Pond, L. D. Van Vleck, and J. K. Loosli. 1965. Effect of ferrous fumarate in the lactation diet on sow milk iron, pig hemoglobin and weight gain. J. Anim. Sci. 24:1169–1173. Vipperman, P. E., Jr., E. R. Peo, Jr., and P. J. Cunningham. 1974. Effect of dietary calcium and phosphorus level upon calcium, phosphorus and nitrogen balance in swine. J. Anim. Sci. 38:758–765. Wahlstrom, R. C., and E. W. Juhl. 1960. A comparison of different methods of iron administration on rate of gain and hemoglobin level of the baby pig. J. Anim. Sci. 19:183–188. Wahlstrom, R. C., and G. W. Libal. 1981. Influence of supplemental dietary potassium on performance of growing-finishing swine. Swine Day Rep. ASR 81-11. Brookings, SD: South Dakota State University. Wahlstrom, R. C., L. D. Kamstra, and O. E. Olson. 1955. The effect of arsanilic acid and 3-nitro-4-hydroxyphenylarsonic acid on selenium poisoning in the pig. J. Anim. Sci. 14:105–110. Wallace, H. D. 1967. High Level Copper in Swine Feeding. New York: International Copper Research Association, Inc. Ward, T. L., L. L. Southern, and R. A. Anderson. 1995. Effect of dietary chromium source on growth, carcass characteristics, and plasma metabolite and hormone concentrations in growing-finishing swine. J. Anim. Sci. 73(Suppl. 1):189 (Abstr.). Ward, T. L., G. L. Asche, G. F. Louis, and D. S. Pollmann. 1996. Zincmethionine improves growth performance of starter pigs. J. Anim. Sci. 74(Suppl. 1):182 (Abstr.). Wedekind, K. J., A. J. Lewis, M. A. Giesemann, and P. S. Miller. 1994. Bioavailability of zinc from inorganic and organic sources for pigs fed corn–soybean meal diets. J. Anim. Sci. 72:2681–2689. Weeden, T. L., J. L. Nelssen, R. D. Goodband, J. A. Hansen, K. G. Fiesen, and B. T. Richert. 1993a. The interrelationship of porcine somatotropin administration and dietary phosphorus on growth performance and bone properties in developing gilts. J. Anim. Sci. 71:2683–2692. Weeden, T. L., J. L. Nelssen, R. D. Goodband, J. A. Hansen, G. E. Fitzner, K. G. Fiesen, and J. L. Laurin. 1993b. Effects of porcine somatotropin and dietary phosphorus on growth performance and bone properties of gilts. J. Anim. Sci. 71:2674–2682. Wegger, I., K. Rasmussen, and P. F. Jorgensen. 1980. Glutathione peroxidase activity in liver and kidney as indicator of selenium status in swine. Livestock Prod. Sci. 7:175–180.

Weinberg, E. D. 1978. Iron and infection. Microbiol. Res. 42:45–66. Welch, H. 1928. Goiter in farm animals. Mont. Agric. Exp. Stn. Bull. 214:1–27. White, M., J. Pettigrew, J. Zollitsch-Stelzl, and B. Crooker. 1993. Chromium in swine diets. Pp. 251–261 in Proc. 54th Minn. Nutr. Conf. and Nat. Renderers Tech. Symp. Whiting, F., and L. M. Bezeau. 1958. The calcium, phosphorus and zinc balance in pigs as influenced by the weight of pig and the level of calcium, zinc and vitamin D in the ration. Can. J. Anim. Sci. 38:109–117. Wilde, R. O. de, and J. Jourquin. 1992. Estimation of digestible phosphorus requirements in growing-finishing pigs by carcass analysis. J. Anim. Phys. Anim. Nutr. 68:218. Wilkinson, J. E., M. C. Bell, J. A. Bacon, and F. B. Masincupp. 1977a. Effects of supplemental selenium on swine. I. Gestation and lactation. J. Anim. Sci. 44:224–228. Wilkinson, J. E., M. C. Bell, J. A. Bacon, and C. C. Melton. 1977b. Effects of supplemental selenium on swine. II. Growing-finishing. J. Anim. Sci.44:229–233. Young, L. G., J. H. Lumsden, A. Lun, J. Claxton, and D. E. Edmeades. 1976. Influence of dietary levels of vitamin E and selenium on tissue and blood parameters in pigs. Can. J. Comp. Med. 40:92–97. Young, L. G., M. Leunissen, and J. L. Atkinson. 1993. Addition of microbial phytase to diets of young pigs. J. Anim. Sci. 71:2147–2151. Zhou, W., E. T. Kornegay, and M. D. Lindemann. 1994a. The role of feed intake and copper source on copper-stimulated growth in weanling pigs. J. Anim. Sci. 72:2385–2394. Zhou, W., E. T. Kornegay, M. D. Lindemann, J. W. G. M. Swinkels, M. K. Welten, and E. A. Wong. 1994b. Stimulation of growth by intravenous injection of copper in weanling pigs. J. Anim. Sci. 72:2395–2043. Zimmerman, D. R. 1980. Iron in swine nutrition. In National Feed Ingredient Association Literature Review on Iron in Animal and Poultry Nutrition. Des Moines, Iowa: National Feed Ingredient Association. Zimmerman, D. R. 1982. Lysine and potassium levels in pig starter diets. J. Anim. Sci. 55(Suppl. 1):97. Zimmerman, D. R., V. C. Speer, V. W. Hays, and D. V. Catron. 1959. Injectable iron dextran and several oral iron treatments for the prevention of iron deficiency anemia of baby pigs. J. Anim. Sci. 18:1409–1415. Zimmerman, D. R., V. C. Speer, V. W. Hays, and D. V. Catron. 1963. Effect of calcium and phosphorus levels on baby pig performance. J. Anim. Sci. 22:658–661. Zoubek, G. L., E. R. Peo, Jr., B. D. Moser, T. Stahly and P. J. Cunningham. 1975. Effects of source on copper uptake by swine. J. Anim. Sci. 40:880–884.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

5

Vitamins

The term ‘‘vitamin’’ describes an organic compound distinct from amino acids, carbohydrates, and lipids that is required in minute amounts for normal growth and reproduction. Some vitamins are not required in the diet because they can be synthesized readily from other feed or metabolic constituents, or by microorganisms in the intestinal tract. Vitamins are generally classified as either fat-soluble or water-soluble. The fat-soluble vitamins include vitamins A, D, E, and K. The water-soluble vitamins include the B-vitamins (biotin, choline, folacin, niacin, pantothenic acid, riboflavin, thiamin, B 6 , and B 12 ) and vitamin C (ascorbic acid). Vitamins are primarily required as coenzymes in nutrient metabolism. In feedstuffs, vitamins exist primarily as precursor compounds or coenzymes that may be bound or complexed in some manner. Hence, digestive processes are required to either release or convert vitamin precursors or complexes to usable and absorbable forms. The requirements for the individual vitamins at various stages of the life cycle are shown in tables provided in Chapter 10. To meet the deficiencies of vitamins in practical diets, vitamin premixes have been developed and are commonly added to swine diets. Dietary addition of excess levels of vitamins A and D to the diet has been demonstrated to have toxic effects in swine. In contrast, very few toxicity signs have been reported for the B-vitamins or for vitamins E and K (National Research Council, 1987). Several recent studies have suggested that National Research Council (1988) levels of one or more of the commonly supplemented B-vitamins (riboflavin, niacin, pantothenic acid, and vitamin B12) are inadequate for maximal performance of newly weaned pigs (Wilson et al., 1991a,b; 1992a,b; 1993) or high-lean growing pigs (Stahly et al., 1995). Indeed, additions of these B-vitamins at levels of two to ten times the estimated requirements have tended to improve growth rate or feed efficiency of pigs. However,

it is not known what level (above those suggested by the National Research Council in 1988) may be needed. Lindemann et al. (1995) observed a trend toward improved weight gain and feed intake in weanling pigs fed five times National Research Council (1988) levels of commonly supplemented vitamins (including fat-soluble vitamins), but feed efficiency tended to be poorer with the higher levels of vitamin fortification. In a separate study, the same group tested a level of vitamin B12 7.5 times the 1988 standard and observed no positive responses. In most of these previously mentioned studies, combinations of vitamins were added and fortification levels were such that it is not possible to establish revised estimates of requirements for individual B-vitamins. Therefore, the B-vitamin requirements for weanling pigs have not been changed. More research certainly is needed to clarify this issue.

FAT-SOLUBLE VITAMINS Vitamin A Vitamin A is essential for vision, reproduction, the growth and maintenance of differentiated epithelia, and mucus secretions. Except for its role in vision (Wald, 1968), the exact role of vitamin A in these functions is undefined (Goodman, 1979, 1980). Recent evidence, however, suggests that vitamin A may be involved in gene expression. Vitamin A nomenclature policy (Anonymous, 1990) dictates that the term ‘‘vitamin A’’ be used for all b-ionone derivatives, other than provitamin A carotenoids, that exhibit the biological activity of all-trans retinol (i.e., vitamin A alcohol, or retinol). Vitamin A is present in animal tissues, eggs, and whole milk, whereas plant materials contain only provitamin A precursors that must be acted upon in the gut or by the liver to form retinol. Both natural vitamin A and synthetic retinol analogs are commonly

71

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

72

Nutrient Requirements of Swine

referred to as retinoids. On the basis of rat data, 1 IU of vitamin A equals 0.3 mg of crystalline vitamin A alcohol, 0.344 mg of vitamin A acetate, or 0.55 mg of vitamin A palmitate. Retinol equivalent (RE) is the currently accepted nomenclature used to describe the vitamin activity in foods and feeds. One RE is defined as 1 mg of alltrans retinol. Pigs are less efficient than poultry or rats in converting carotenoid precursors to vitamin A. This conversion occurs primarily in intestinal mucosa (Fidge et al., 1969). Active carotenoid pigments in corn–soybean meal diets (Wellenreiter et al., 1969) and their bioactivities relative to alltrans b-carotene (100 percent) are b-zeacarotene (25 percent) and cryptoxanthin (57 percent), as estimated by Petzold et al. (1959), Duel et al. (1945), and Greenberg et al. (1950). Ullrey (1972) calculated, therefore, that the alltrans b-carotene equivalent would be only 52 percent of the chemically determined carotene value. He then calculated that this value for swine would be only 16 percent, based on the fact that pigs are only 30 percent as efficient as rats in converting b-carotene in swine diets to usable vitamin A (Braude et al., 1941). When this value is multiplied by 1,667 IU, which represents the theoretical vitamin A potency of 1 mg of all-trans b-carotene for rats, 1 mg of chemically determined carotene in a corn–soybean meal pig diet would have a calculated potency of 267 IU, or 80 mg of vitamin A alcohol. Chew et al. (1982) and Brief and Chew (1985) have suggested that b-carotene plays a role in reproduction that is independent of vitamin A. Their studies involving bcarotene injection suggest that elevation of maternal plasma vitamin A or b-carotene may improve embryonic survival, possibly because more uterine-specific proteins are secreted. Dietary addition of b-carotene did not elicit a response. This failure is probably due to the poor absorption of intact b-carotene in the pig (Poor et al., 1987). Swine are able to store vitamin A in the liver, which makes the vitamin available during periods of low intake. Requirements for vitamin A depend on the criteria evaluated; weight gain is less sensitive than cerebrospinal fluid pressure, liver storage, or plasma levels. For pigs during the first 8 weeks of life, 75 to 605 mg of retinyl acetate/kg of diet is required, depending on the response criteria used (Sheffy et al., 1954; Frape et al., 1959). With growingfinishing pigs, the requirement varies from 35 to 130 mg/kg, when daily gain is used as the criterion, and from 344 to 930 mg/kg, when liver storage and cerebrospinal fluid pressure are used as the criteria (Guilbert et al., 1937; Braude et al., 1941; Hentges et al., 1952a; Myers et al., 1959; Hjarde et al., 1961; Nelson et al., 1962; Ullrey et al., 1965). Presence of nitrite or nitrate in feed or water can increase the vitamin A requirement (Koch et al., 1963; Seerley et al., 1965; Wood et al., 1967; Hutagalung et al., 1968).

The vitamin A reserves of the sow make it difficult to establish requirements. Braude et al. (1941) reported that mature sows fed diets without supplemental vitamin A completed three pregnancies normally; only in the fourth pregnancy did signs of vitamin deficiency appear. Gilts receiving adequate vitamin A levels until 9 months of age, followed by a diet containing no vitamin A, completed two reproductive cycles without signs of vitamin A deficiencies (Hjarde et al., 1961; Selke et al., 1967). Heaney et al. (1963) fed depleted gilts 16, 5, or 2.5 mg of retinyl palmitate/kg body weight daily with no effects on litter size, birth weight, or survival rate. Parrish et al. (1951) suggested that 2,100 IU of vitamin A/day during gestation and lactation was adequate to maintain normal serum and liver concentrations. Vitamin A deficiency in swine results in reduced weight gain, incoordination, posterior paralysis, blindness, increased cerebrospinal fluid pressure, decreased plasma levels, and reduced liver storage (Guilbert et al., 1937; Braude et al., 1941; Hentges et al., 1952a; Frape et al., 1959; Hjarde et al., 1961; Nelson et al., 1962, 1964). Gross toxicity signs of hypervitaminosis A include a roughened hair coat, scaly skin, hyperirritability and sensitivity to touch, bleeding from cracks which appear in the skin about the hooves, blood in urine and feces, loss of control of the legs accompanied by inability to rise, and periodic tremors (Anderson et al., 1966). Young pigs fed diets containing 605,000, 484,000, 363,000, or 242,000 mg of retinyl palmitate/kg of diet developed signs of vitamin A toxicity in 16, 17.5, 32, and 43 days, respectively. No signs of toxicity were observed when pigs were fed 121,000 mg of added retinyl palmitate/kg of diet for 8 weeks (Anderson et al., 1966). Wolke et al. (1968) observed lesions in endochondral and intramembranous bone within 5 weeks when pigs were fed these excessive levels of vitamin A. Vitamin A esters are more stable in feeds and premixes than is retinol. The hydroxyl group as well as the four double bonds on the retinol side chain are subject to oxidative losses. Thus, esterification of vitamin A alcohol does not totally protect this vitamin from oxidative losses. Current commercial sources of vitamin A are generally ‘‘coated’’ esters (1 IU of vitamin A 4 0.344 mg of retinyl acetate, or 0.549 mg of retinyl palmitate) that contain an added antioxidant such as ethoxyquin or butylated hydroxytoluene (BHT). Moisture in premixes and feedstuffs has a negative effect on vitamin A stability (Baker, 1995). Water causes vitamin A beadlets to soften and become more permeable to oxygen. Thus, both high humidity and presence of free choline chloride (which is very hygroscopic) enhance vitamin A destruction. Trace minerals also exacerbate vitamin A losses in premixes exposed to moisture. For maximum retention of vitamin A activity, premixes should be as moisture-free as possible and have a pH above five. Low pH

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins causes isomerization of all-trans vitamin A to less potent cis forms and also results in de-esterification of vitamin A esters to more labile retinol (De Ritter, 1976).

Vitamin D The two major forms of vitamin D are ergocalciferol (vitamin D2) and cholecalciferol (vitamin D3). The action of ultraviolet light on the ergosterol that is present in plants forms ergocalciferol; the photochemical conversion of 7dehydrocholesterol in the skin of animals forms cholecalciferol. One IU of vitamin D is defined as the biological activity of 0.025 mg of cholecalciferol. Ergocalciferol and cholecalciferol are hydroxylated in the liver to the 25hydroxy forms. The 25-hydroxy-D3 is further hydroxylated in the kidney to either 1,25-dihydroxy-D3 or 24,25-dihydroxy-D3. Several mechanisms that act according to established criteria for hormones control the synthesis and reactions of the dihydroxylated metabolites; therefore, the dihydroxylated D 3 metabolites are viewed as hormones (Schnoes and DeLuca, 1980; Kormann and Weiser, 1984). Vitamin D and its hormonal metabolites act on the mucosal cells of the small intestine, causing the formation of calcium-binding proteins. These proteins facilitate calcium and magnesium absorption and influence phosphorus absorption. The actions of vitamin D metabolites, together with parathyroid hormone and calcitonin, maintain calcium and phosphorus homeostasis. Braidman and Anderson (1985) have reviewed the endocrine functions of vitamin D. Bethke et al. (1946) indicated that vitamins D2 and D3 were equally effective in meeting the vitamin D needs of swine. Horst et al. (1982), however, demonstrated that pigs discriminate in their metabolism of the two forms of vitamin D. Additional research is needed in swine to quantify the differences in absorption and utilization of these forms. The vitamin D 2 requirement of the baby pig fed a casein–glucose diet is 100 IU/kg of diet (Miller et al., 1964, 1965). The requirement is higher if isolated soy protein is fed (Miller et al., 1965; Hendricks et al., 1967). Vitamin D deficiency reduces retention of calcium, phosphorus, and magnesium (Miller et al., 1965). Bethke et al. (1946) suggested a minimum requirement of 200 IU/kg of diet for growing pigs. In other studies, however, vitamin D supplementation did not improve weight gain (Wahlstrom and Stolte, 1958; Combs et al., 1966). No studies of the vitamin D requirement of sows during gestation or lactation have been reported. Weisman et al. (1976), Boass et al. (1977), Noff and Edelstein (1978), Halloran and DeLuca (1979), and Pike et al. (1979) showed that vitamin D is involved in rat reproduction and lactation. Parenteral cholecalciferol treatment of sows before parturition provided an effective means of supplementing pigs

73

with cholecalciferol (via the sow’s milk) and its dihydroxy metabolites by placental transport (Goff et al., 1984). Vitamin D deficiency causes a disturbance in the absorption and metabolism of calcium and phosphorus that results in insufficient bone calcification. In young growing pigs, vitamin D deficiency results in rickets, whereas in mature swine a deficiency causes diminished bone mineral content (osteomalacia). In severe vitamin D deficiency, pigs may exhibit signs of calcium and magnesium deficiency, including tetany. It takes 4 to 6 months for pigs fed a vitamin D–deficient diet to develop signs of a deficiency (Johnson and Palmer, 1939; Quarterman et al., 1964). Vitamin D toxicity was produced in weanling pigs supplemented with a daily oral dose of 6,250 mg of vitamin D3 for 4 weeks (Quarterman et al., 1964). This level of D3 reduced feed intake; growth rate; and weights of the liver, radius, and ulna. At necropsy, calcification was observed in the aorta, heart, kidney, and lung. Feeding a daily level of 11,825 mg of vitamin D3 to pigs weighing 20 to 25 kg resulted in death in 4 days (Long, 1984). Hancock et al. (1986) reported that there was a reduction in daily gain and feed efficiency in pigs weighing 10 to 20 kg fed a diet containing 550 to 1,100 mg/kg of supplemental vitamin D3/ day. Vitamin D3 has been shown to be more toxic than vitamin D 2 in a number of species, including swine (National Research Council, 1987). The development of methods to measure vitamin D and its metabolites in plasma has provided insights regarding the possible mechanisms that cause differences in toxicity between vitamins D2 and D3 (Horst et al., 1981; National Research Council, 1987). Vitamin E There are eight naturally occurring forms of vitamin E: a, b, g, and d tocopherols (Evans et al., 1936; Emerson et al., 1937; Stern et al., 1947) and a, b, g, and d tocotrienols (Green et al., 1960; Pennock et al., 1964; Whittle et al., 1966). Of these, D-a–tocopherol possesses the greatest biological activity (Brubacher and Wiss, 1972; Ames, 1979; Bieri and McKenna, 1981). One IU of vitamin E is the activity of 1 mg of DL-a–tocopheryl acetate. The D isomer is more bioactive than the L isomer. On the basis principally of rat bioassay work and using DL-a–tocopheryl acetate as a standard (1 mg 4 1 IU), it is calculated that 1 mg DLa–tocopherol equals 1.1 IU, 1 mg D-a–tocopheryl acetate equals 1.36 IU, and 1 mg D-a–tocopherol equals 1.49 IU of vitamin E. Anderson et al. (1995a), however, suggested that D-a–tocopheryl acetate is utilized more efficiently by pigs than by rats. For young pigs, Chung et al. (1992) reported that 1 mg D-a–tocopherol equals 2.44 IU. For many years the primary source of vitamin E in feed was the tocopherols found in green plants and seeds. Oxidation, which is accelerated by heat, moisture, rancid

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

74

Nutrient Requirements of Swine

fat, and trace minerals, rapidly destroys natural vitamin E. Therefore, predicting the amount of vitamin E activity in feed ingredients is difficult. Vitamin E losses of 50 to 70 percent can occur in alfalfa stored at 32°C for 12 weeks; losses of 5 to 30 percent can occur during dehydration of alfalfa (Livingston et al., 1968). Storage of high-moisture grain or its treatment with organic acids greatly reduces its vitamin E content (Madsen et al., 1973; Lynch et al., 1975; Young et al., 1975, 1978). High levels of dietary vitamin A have also been reported to lower vitamin E absorption (Hoppe et al., 1992), although Anderson et al. (1995b) observed no effects on vitamin E status when growing pigs were fed diets containing 15 times the vitamin A requirement. During the 1970s, many studies on the vitamin E requirement of swine were conducted. The Agricultural Research Council (1981) and Ullrey (1981) have reviewed the studies. Many dietary factors affect the vitamin E requirement, including levels of selenium, unsaturated fatty acids, sulfur amino acids, retinol, copper, iron, and synthetic antioxidants. Michel et al. (1969) prevented deaths in pigs fed a corn–soybean diet containing 5 to 8 mg of vitamin E/kg and 0.04 to 0.06 mg of selenium/kg by supplementing the diet with 22 mg of vitamin E/kg. Studies of corn–soybean meal diets fed to growing-finishing pigs suggest that 5 mg of vitamin E/kg and 0.04 mg of selenium/ kg are inadequate for growing-finishing pigs and may result in deficiency lesions and mortality. In the presence of adequate selenium, however, supplements of 10 to 15 mg of vitamin E/kg of diet prevented mortality and deficiency lesions and supported normal performance (Groce et al., 1971, 1973; Sharp et al., 1972a,b; Ullrey, 1974; Wilkinson et al., 1977b; Hitchcock et al., 1978; Mahan and Moxon, 1978; Meyer et al., 1981). The amount of vitamin E necessary to prevent deficiency signs varies considerably because of variation in dietary levels of selenium (Agricultural Research Council, 1981; Ullrey, 1981), antioxidants (Tollerz, 1973; Simesen et al., 1982), and lipids (Nielsen et al., 1973; Tiege et al., 1977, 1978). Inclusion of high levels of vitamin E in the diet may increase the immune response (Ellis and Vorhies, 1976; Tiege, 1977; Nockels, 1979; Peplowski et al., 1980; Wuryastuti et al., 1993), although Bonnette et al. (1990) found no evidence of an increased humoral or cell-mediated immune response in young pigs fed high levels of vitamin E. Vitamin E functions as an antioxidant at the cell membrane level, and it has a structural role in cell membranes. There are vitamin E deficiency diseases that respond to vitamin E, selenium, or antioxidants. Vitamin E deficiency results in a wide variety of pathological conditions. These include skeletal and cardiac muscle degeneration, degenerative thrombotic vessel injury, gastric parakeratosis, gastric ulcers, anemia, liver necrosis, yellow discoloration of fat tissue, and sudden death (Obel, 1953; Davis and Gorham,

1954; Hove and Seibold, 1955; Dodd and Newling, 1960; Grant, 1961; Lannek et al., 1961; Nafstad, 1965, 1973; Nafstad and Nafstad, 1968; Reid et al., 1968; Ewan et al., 1969; Michel et al., 1969; Nafstad and Tollersrud, 1970; Trapp et al., 1970; Baustad and Nafstad, 1972; Sharp et al., 1972a,b; Sweeney and Brown, 1972; Wastell et al., 1972; Piper et al., 1975; Bengtsson et al., 1978a,b; Hakkarainen et al., 1978; Tiege and Nafstad, 1978; Simesen et al., 1982). In addition, vitamin E may be involved in the mastitismetritis-agalactia (MMA) complex of sows (Ringarp, 1960; Ullrey et al., 1971; Whitehair et al., 1984). Information is available on the vitamin E requirements for reproduction (Hanson and Hathaway, 1948; Adamstone et al., 1949; Cline et al., 1974; Malm et al., 1976; Young et al., 1977, 1978; Wilkinson et al., 1977a; Nielsen et al., 1979; Piatkowski et al., 1979; Mahan, 1991, 1994). Placental transfer of tocopherol from dam to fetus is minimal, so the offspring must rely on colostrum and milk to meet their daily needs. The content of vitamin E in sow colostrum and milk is dependent on the vitamin E content of the sow’s diet (Mahan, 1991). In most studies, diets containing 5 to 7 mg/kg of vitamin E and 0.1 mg/kg of inorganic selenium have prevented deficiency lesions and supported normal reproductive performance. But the addition of 0.1 mg/kg of inorganic selenium and 22 mg/kg of vitamin E to diets appears necessary to maintain tissue vitamin E levels (Piatkowski et al., 1979). Recent research (Mahan, 1991; 1994; Wuryastuti et al., 1993), however, suggests that vitamin E levels as high as 44 to 60 mg/kg during gestation and lactation may be necessary to maximize both litter size and immunocompetency. As a result of these recent findings, the vitamin E requirements for gestation and lactation have been increased to 44 IU/kg of diet. Vitamin E toxicity has not been demonstrated in swine. Levels as high as 550 mg/kg of diet have been fed to growing pigs without toxic effects (Bonnette et al., 1990). Vitamin K Although it was the last of the four fat-soluble vitamins to be discovered, the metabolic role of vitamin K has been more clearly defined than that of vitamins A, D, and E (Suttie, 1980; Kormann and Weiser, 1984). Vitamin K is essential for the synthesis of prothrombin, factor VII, factor IX, and factor X, which are necessary for the normal clotting of blood. These proteins are synthesized in the liver as inactive precursors. The action of vitamin K converts them to biologically active compounds (Suttie and Jackson, 1977; Suttie, 1980). This activation occurs by enzymatic g-carboxylation of specific glutamate residues. The resulting carboxyglutamate residues are strong chelators of calcium ions, which are essential for blood coagulation. A deficiency of vitamin K or the presence of anticoagulation compounds reduces the number of carboxyglutamate residues, result-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins ing in a loss of activity and prolonged bleeding times. In addition to its role in blood clotting, there is evidence that vitamin K–dependent protein and peptides may be involved in calcium metabolism (Suttie, 1980; Kormann and Weiser, 1984). Vitamin K exists in three series: the phylloquinones (K1) in plants; the menaquinones (K2), formed by microbial fermentation; and the menadiones (K3), which are synthetic. Menadione (2-methyl-1,4-naphthoquinone) is the synthetic form of vitamin K, which has the same cyclic structure as vitamins K1 and K2. All three forms of vitamin K are biologically active. Water-soluble forms of menadione are commonly used to supplement swine diets. The major forms are menadione sodium bisulfite (MSB) and menadione dimethylpyrimidinol bisulfite (MPB). Menadione sodium bisulfite complex (MSBC) is used in poultry diets, but it does not have FDA approval for use in swine diets. The vitamin K activity depends upon the menadione content of these products: 52, 33, and 46 percent menadione in MSB, MSBC, and MPB, respectively. Menadione nicotinamide bisulfite is a new synthetic form of vitamin K that has been shown to have a bioactivity similar to that of MPB (Oduho et al., 1993). Vitamin K deficiency increases prothrombin and clotting times and may result in internal hemorrhages and death (Schendel and Johnson, 1962; Brooks et al., 1973; Seerley et al., 1976; Hall et al., 1986, 1991). Schendel and Johnson (1962) reported a requirement of 5 mg of menadione sodium phosphate/kg of body weight for 1- and 2-dayold pigs fed a purified liquid diet. Their diet contained sulfathiazole and oxytetracycline to reduce the intestinal synthesis of vitamin K. Wire-bottomed cages were used and carefully cleaned to minimize coprophagy. Seerley et al. (1976) reported that 1.1 mg of MPB/kg of diet counteracted the effects of the anticoagulant pivalyl in weanling pigs. Hall et al. (1986) suggested that 2 mg/kg of menadione as MPB was needed to counteract the effects of pivalyl in growing pigs. Bacterial synthesis of vitamin K and subsequent absorption, directly or by coprophagy, reduces or eliminates the need for supplemental vitamin K. High levels of antibiotics may decrease the synthesis of vitamin K by the intestinal flora. Studies have not been conducted to determine whether a supplemental source of vitamin K is beneficial for the breeding herd. Muhrer et al. (1970), Osweiler (1970), and Fritschen et al. (1971) reported an occurrence of hemorrhagic conditions in pigs under field conditions. Mycotoxin-contaminated ingredients were suspected in these incidents, and vitamin K supplementation (2.0 mg of menadione/kg of diet) prevented the hemorrhagic syndrome. In some of these studies, the presence of anti-clotting coumarins may have increased the dietary requirement for vitamin K.

75

Excess calcium may also increase the pig’s requirement for vitamin K (Hall et al., 1991). Liver stores of vitamin K can be depleted very rapidly during even very short periods of vitamin K–deficient diet consumption (Kindberg and Suttie, 1989). Stability of water-soluble menadione supplements in premixes and diets is impaired by moisture, choline chloride, trace elements, and alkaline conditions. Coelho (1991) suggested that MSBC and MPB can lose up to 80 percent of bioactivity if stored for 3 months in a vitamin–trace– mineral premix containing choline. Activity losses were far less when the menadione compounds were stored in the same premix that did not contain choline. Some menadione supplements are now coated, and this appears to improve stability in diets and premixes. Even very large amounts of menadione compounds are tolerated well by animals. Seerley et al. (1976) fed 110 mg/ kg of MPB to pigs, and Oduho et al. (1993) fed 300 mg/ kg of MPB to chicks; neither observed signs of toxicity. A dietary level of 3,000 mg/kg of MPB did not depress weight gain or blood hemoglobin when fed over a 14-day period to chicks. It appears that menadione levels of 1,000 times an animal’s requirement are well tolerated (National Research Council, 1987; Oduho et al., 1993).

WATER-SOLUBLE VITAMINS Biotin Biotin is important metabolically as a cofactor for several enzymes that function in carbon dioxide fixation. As part of pyruvate carboxylase and propionyl CoA carboxylase, it is important in gluconeogenesis and in the citric acid cycle. Acetyl CoA carboxylase is also a biotin-dependent enzyme that functions in initiating fatty acid biosynthesis. Whitehead et al. (1980) and Misir and Blair (1986) suggested that plasma biotin concentration and plasma pyruvate carboxylase activity are methods of assessing the biotin status of pigs. The D-isomer of biotin is the biologically active form of the vitamin. Biotin is present in most common feedstuffs in morethan-adequate amounts, but its bioavailability varies greatly among ingredients. The bioavailability of biotin in yelllow corn and soybean meal is high for the chick, but its bioavailability in barley, grain sorghum, oats, and wheat is lower (Frigg, 1976; Anderson et al., 1978; Kopinski et al., 1989). Much of the biotin in feed ingredients exists as e-N-biotinyl L-lysine (biocytin), which is a component of protein. The bioavailability of biocytin (relative to crystalline D-biotin) varies widely and is dependent on the digestibility of the proteins in which it is found. A considerable portion of the pig’s biotin requirement is presumed to come from bacterial synthesis in the gut.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

76

Nutrient Requirements of Swine

In general, performance has not been improved by supplemental biotin in a wide range of diets and conditions for pigs weaned at 2 to 28 days of age or for growingfinishing pigs. Pigs from 2 to 28 days of age fed a filtered skim milk diet containing about 10 mg of biotin/kg of dry matter (about 15 percent of the level in sow’s milk) gained weight and were as efficient in feed conversion as littermate pigs supplemented with 50 mg of biotin/kg of diet (Newport, 1981). Likewise, biotin supplementation at levels varying from 110 to 880 mg/kg of diet yielded no improvement in rate or efficiency of gain in pigs weaned at 21 to 28 days of age or in growing-finishing pigs (Peo et al., 1970; Hanke and Meade, 1971; Meade, 1971; Washam et al., 1975; Simmins and Brooks, 1980; Easter et al., 1983; Bryant et al., 1985b; Hamilton and Veum, 1986). Exceptions include one experiment that Adams et al. (1967) reported for growing pigs and one experiment that Peo et al. (1970) reported for pigs weaned at 28 days of age. Also, Partridge and McDonald (1990) observed feed efficiency responses to biotin when it was added to wheat–barley–soybean meal diets for growing pigs. With sows, biotin supplementation has been reported to improve hoof hardness and compression, compressive strength, and the condition of skin and hair coat, as well as to reduce hoof cracks and footpad lesions (Grandhi and Strain, 1980; Webb et al., 1984; Bryant et al., 1985a,b; Simmins and Brooks, 1985; Misir and Blair, 1986). However, in studies by Hamilton and Veum (1984) and Tribble et al. (1984), no such improvements were recorded. Lewis et al. (1991) reported that adding 0.33 mg/kg of biotin to a corn–soybean meal diet for sows during both gestation and lactation increased the number of pigs weaned but did not improve foot health. Watkins et al. (1991) also conducted a large-scale biotin efficacy trial for sows during gestation and lactation and reported that none of the criteria of reproductive performance, progeny development, or foot health responded to 0.44 mg of supplemental biotin/kg of diet. Other studies by investigators using a variety of grain sources have resulted in inconsistent results (Brooks et al., 1977; Penny et al., 1981; Easter et al., 1983; Simmins and Brooks, 1983; Hamilton and Veum, 1984; Tribble et al., 1984; Bryant et al., 1985c; Kornegay, 1986; Misir and Blair, 1986). A lack of consistency among experiments and a wide range of biotin supplementation levels (0.1 to 0.55 mg/kg of diet) make it difficult to establish a specific biotin requirement for sows. Biotin deficiency signs include excessive hair loss, skin ulcerations and dermatitis, exudate around the eyes, inflammation of the mucous membranes of the mouth, transverse cracking of the hooves, and the cracking or bleeding of the footpads (Cunha et al., 1946; 1948; Lindley and Cunha, 1946; Lehrer et al., 1952). Biotin deficiency in pigs has been produced by feeding pigs synthetic diets containing sulfa drugs, which presumably reduce the syn-

thesis of biotin in the intestinal tract (Lindley and Cunha, 1946; Cunha et al., 1948; Lehrer et al., 1952). Incorporation of large amounts of desiccated eggwhite in synthetic diets also has precipitated biotin deficiency in pigs (Cunha et al., 1946; Hamilton et al., 1983). Avidin, contained in raw eggwhite, forms a complex with biotin in the intestinal tract, rendering the vitamin unavailable to the pig. Choline Choline remains in the B-vitamin category even though the quantity required far exceeds the ‘‘trace organic nutrient’’ definition of a vitamin. It is generally added to swine diets as choline chloride, which contains 74.6 percent choline activity (Emmert et al., 1996). Choline is required for (a) phospholipid (i.e., lecithin) synthesis, (b) acetyl choline formation, and (c) transmethylation of homocysteine to methionine, which occurs via betaine, the oxidation product of choline. When severe choline deficiency is encountered, phospholipid and acetyl choline synthesis take priority over choline’s methylation functions; however, grain–oilseed meal diets contain enough choline such that betaine or choline is equally efficacious on a molar basis in meeting the methylation function of choline (Lowry et al., 1987). Pigs synthesize choline by methylating phosphatidyl ethanolamine in a three-step process involving methyl transfer from S-adenosylmethionine. Thus, excess dietary methionine can eliminate the dietary need for choline in pigs (Neumann et al., 1949; Nesheim and Johnson, 1950; Kroening and Pond, 1967). Choline from soybean meal has been estimated to be 65 to 83 percent bioavailable relative to choline from choline chloride (Molitoris and Baker, 1976; Emmert and Baker, 1997). Analytical and bioavailability studies with chicks have indicated that dehulled soybean meal contains 2,218 mg of total choline/kg and 1,855 mg of bioavailable choline/ kg; bioavailability of choline in peanut meal (71 percent) was slightly less than that in soybean meal (83 percent) and the choline in canola meal was only 24 percent bioavailable (Emmert and Baker, 1997). Because soy products are rich in bioavailable choline, starting, growing, and finishing pigs have not shown responses to supplemental choline when it was added to corn–soybean meal or corn-isolated soy protein diets (Bryant et al., 1977; Russett et al., 1979b; North Central Region-42 Committee on Swine Nutrition, 1980). A portion of the choline present in feed ingredients and unprocessed fat sources exists as phospholipid-bound choline. This form of choline is thought to be utilized well (Emmert et al., 1996), but refined oils have been subjected to degumming, and this process removes virtually all of the phospholipid-bound choline (Anderson et al., 1979). Feeding pregnant gilts and sows grain–soybean meal diets supplemented with 434 to 880 mg of choline/kg has

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins generally increased the number of live pigs born and weaned (Kornegay and Meacham, 1973; Stockland and Blaylock, 1974; North Central Region-42 Committee on Swine Nutrition, 1976; Grandhi and Strain, 1980). In a long-term reproduction study, Stockland and Blaylock (1974) also reported that choline supplementation of corn– soybean meal diets improved conception rate. Gilts fed a choline-supplemented diet during gestation farrowed heavier pigs, but the incidence of spraddle-legged pigs was not reduced in four trials reported by Luce et al. (1985). During lactation, choline supplementation of diets containing 8 to 10 percent fat or oil did not improve lactation performance (Seerley et al., 1981; Boyd et al., 1982). Choline-deficient pigs have reduced weight gain, rough hair coats, decreased red blood cell counts and hematocrit and hemoglobin concentrations, increased plasma alkaline phosphatase, and unbalanced and staggering gaits. Livers and kidneys exhibit fat infiltration. In a severe choline deficiency, kidney glomeruli can become occluded from massive fat infiltration (Wintrobe et al., 1942; Johnson and James, 1948; Neumann et al., 1949; Russett et al., 1979b). The addition of 260 mg of choline/kg to a diet consisting of 30 percent vitamin-free casein, 37 percent glucose, 26.6 percent lard, and 2 percent sulfathaladine, which contained 0.8 percent methionine, prevented a choline deficiency in neonatal pigs (Johnson and James, 1948). A level of 1,000 mg of choline/kg of diet solids optimized weight gain and feed efficiency and prevented fat infiltration of the liver and kidneys in 2-day-old pigs (Neumann et al., 1949). Further addition of 0.8 percent DL-methionine to this diet did not improve the performance of pair-fed pigs supplemented with 1,000 mg of choline/kg of diet (Nesheim and Johnson, 1950). Kroening and Pond (1967) fed 5-kg pigs a low-protein (12 percent) diet supplemented with three levels of DL-methionine: 0, 0.11, or 0.22 percent. The addition of 1,646 mg of choline/kg of diet tended to improve the weight gains and feed conversion of pigs fed the two lower levels of methionine but not those of pigs fed the diet containing 0.22 percent supplemental methionine. Russett et al. (1979a,b) reported a minimum choline requirement of 330 mg/kg of diet for 6- to 14-kg pigs fed a semisynthetic diet containing 0.31 percent methionine and 0.33 percent cystine. No signs of choline toxicity have been reported in swine (National Research Council, 1987), but daily gain reductions have been observed in pigs fed diets containing 2,000 mg/kg of added choline during the starting, growing, and finishing stages (Southern et al., 1986). In another study (Emmert 1997), a dietary choline level of 10,000 mg/kg did not depress growth in 10-kg pigs, nor did a similar level of betaine. Folacin Folacin includes a group of compounds with folic acid activity. Chemically, folacin consists of a pteridine ring,

77

paraaminobenzoic acid (PABA), and glutamic acid. Animal cells cannot synthesize PABA, nor can they attach glutamic acid to pteroic acid. A deficiency of folacin causes a disturbance in the metabolism of single-carbon compounds, including the synthesis of methyl groups, serine, purines, and thymine. Folacin is involved in the conversion of serine to glycine and homocysteine to methionine. The folacin present in feedstuffs exists primarily as a polyglutamate conjugate containing a g-linked polypeptide chain of seven glutamic acid residues. A group of intestinal enzymes known as conjugases (folyl polyglutamate hydrolases) remove all but the last glutamate residue. Only the monoglutamyl form is thought to be absorbed into the intestinal enterocyte. Most of the folacin taken up by the intestinal brush border is reduced to tetrahydrofolic acid (FH4) and then methylated to 5N-methyl FH4. Like thiamin, folacin has a free amino group (on the pteridine ring), and this makes it heat labile, particularly in diets containing reducing sugars such as dextrose or lactose. Except for the studies of Matte et al. (1984a,b; 1992) and Lindemann and Kornegay (1986; 1989), results have indicated that the folacin contribution of ingredients commonly fed to swine when combined with bacterial synthesis within the intestinal tract adequately meets the requirement for all classes of swine. Supplementation of a corn–soybean meal diet with 200 mg of folic acid/kg of diet during pregnancy did not increase the number of pigs born alive or weaned (Easter et al., 1983). Matte et al. (1984a) administered 15 mg of folic acid intramuscularly to sows 10 times, beginning at weaning and continuing until day 60 of pregnancy. They reported a significant increase in litter size farrowed. In a subsequent study, Matte et al. (1992) observed an increase in litter growth rate when the gestation diet was supplemented with 5 or 15 mg/kg of folic acid. Supplementation of the lactation diet, however, did not improve performance of the offspring. Lindemann and Kornegay (1989) also observed increased litter size at birth, but not at weaning, when the corn–soybean meal diet fed to sows was supplemented with 1 mg/kg of folacin. In a study by Tremblay et al. (1986), 4.3 mg of supplemental folic acid/kg of diet (diet containing 0.62 mg of folic acid/kg) maintained serum folate concentrations equivalent to those of pregnant sows injected with folic acid at various intervals from weaning to 56 days after mating (10 injections of 15 mg/sow). In a large multiparity study involving 393 sows, addition of 1, 2, or 4 mg/kg of folic acid to standard corn–soybean meal diets during premating, gestation, and lactation had no beneficial effects on reproductive performance (Harper et al., 1994). Based on these recent studies, the folacin requirement for gestating and lactating sows was increased to 1.3 mg/kg of diet. Folacin deficiency in pigs leads to slow weight gain, fading hair color, macrocytic or normocytic anemia, leuko-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

78

Nutrient Requirements of Swine

penia, thrombopenia, reduced hematocrit, and bone marrow hyperplasia. Synthetic diets, generally with the inclusion of 1 to 2 percent sulfa drugs or folic acid antagonists, have been fed to produce folacin deficiency in pigs (Cunha et al., 1948; Heinle et al., 1948; Cartwright et al., 1949, 1950; Johnson et al., 1950). Sulfa drugs presumably reduce bacterial synthesis of folacin in the intestinal tract. Folic acid supplementation did not affect the performance of 4-day-old pigs fed a synthetic diet that included 2 percent sulfathaladine (Johnson et al., 1948) or of 8-week-old pigs fed a synthetic diet (Cunha et al., 1947). Newcomb and Allee (1986) reported no beneficial effects from the addition of 1.1 mg of folic acid/kg to a corn–soybean meal–whey diet for pigs weaned at 17 to 27 days of age. But Lindemann and Kornegay (1986) observed an improved daily weight gain in pigs of similar age fed a corn–soybean meal diet supplemented with 0.5 mg of folic acid/kg of diet. Pigs fed corn–soybean meal diets during the starting, growing, and finishing phases gained weight and used their feed as efficiently as those supplemented with 200 or 360 mg of folic acid/kg of diet (Easter et al., 1983; Gannon and Liebholz, 1989). Niacin Niacin or nicotinic acid is a component of the coenzymes nicotinamide-adenine dinucleotide (NAD) and nicotinamide-adenine dinucleotide phosphate (NADP). These coenzymes are essential for the metabolism of carbohydrates, proteins, and lipids. Metabolic conversion of excess dietary tryptophan to niacin has complicated the determination of the niacin requirement (Luecke et al., 1948; Powick et al., 1948). Firth and Johnson (1956) estimated that each 50 mg of tryptophan in excess of the tryptophan requirement yields 1 mg of niacin. Niacin status is further complicated by its limited bioavailability in certain feed ingredients. The niacin in yellow corn, oats, wheat, and grain sorghum is in a bound form that is largely unavailable to young pigs (Kodicek et al., 1956; Luce et al., 1966, 1967; Harmon et al., 1969, 1970). The niacin in soybean meal, however, is highly available for the chick and is probably equally available for the pig (Yen et al., 1977). Niacin activity is commercially available as either free nicotinic acid or free nicotinamide (niacinamide). Relative to nicotinic acid, nicotinamide is 124 percent bioavailable for chicks (Oduho and Baker, 1993) and 109 percent bioavailable for rats (Carter and Carpenter, 1982). Firth and Johnson (1956) estimated the available niacin requirements for 1- to 8-kg pigs to be about 20 mg/kg for a diet with no excess tryptophan. Requirement estimates for growing pigs weighing 10 to 50 kg are 10 to 15 mg of available niacin/kg for diets containing tryptophan levels near the requirement (Braude et al., 1946; Kodicek et al.,

1959; Harmon et al., 1969). Growing-finishing diets are usually fortified with niacin, but studies with 45-kg pigs fed corn–soybean meal diets have indicated no performance improvements due to niacin supplementation (Yen et al., 1978; Copelin et al., 1980). The diets used in these experiments, however, contained calculated tryptophan levels that were in excess of the requirement. There is no information on the niacin requirement of pregnant and lactating sows. Research with chicks has demonstrated that iron deficiency impairs the efficacy of tryptophan as a niacin precursor (Oduho et al., 1994). Whether this relationship occurs in pigs is unknown. Iron is required as a cofactor for two enzymes in the pathway leading to nicotinic acid mononucleotide synthesis from tryptophan. Niacin deficiency signs include reduced weight gain, anorexia, vomiting, dry skin, dermatitis, rough hair coat, hair loss, diarrhea, mucosal ulcerations, ulcerative gastritis, inflammation and necrosis of the cecum and colon, and normocytic anemia (Huges, 1943; Wintrobe et al., 1946; Braude et al., 1946; Powick et al., 1947a,b; Luecke et al., 1947; Cartwright et al., 1948; Burroughs et al., 1950; Kodicek et al., 1956). Blood erythrocyte NAD activity and urinary excretions of N-methyl-nicotinamide and N8methyl-2-pyridone-5-carboxamide are reduced in niacin deficiency (Luce et al., 1966, 1967). Pantothenic Acid This B-vitamin consists of pantoic acid joined to b-alanine by an amide bond. As a component of coenzyme A, pantothenic acid is important in the catabolism and synthesis of two-carbon units evolved during carbohydrate and fat metabolism. Biological availability of pantothenic acid is low in barley, wheat, and sorghum but is high in corn and soybean meal (Southern and Baker, 1981). In feedstuffs, most of the pantothenic acid exists as coenzyme A, acyl CoA synthetase, and acyl carrier protein. Only the D-isomer of pantothenic acid is biologically active. Synthetic pantothenic acid is generally added to all swine diets as calcium pantothenate, a salt that is more stable than pantothenic acid. The D-form of calcium pantothenate has 92 percent activity; the racemic mixture of the calcium salt contains only 46 percent active pantothenic acid. A DLcalcium pantothenate–calcium chloride complex is also available, and it contains 32 percent activity. The pantothenic acid requirement of 2- to 10-kg pigs fed synthetic diets was 15.0 mg/kg (Stothers et al., 1955); and for 5- to 50-kg pigs, estimates range from about 4.0 to 9.0 mg/kg of diet (Luecke et al., 1953; Barnhart et al., 1957; Sewell et al., 1962; Palm et al., 1968). Requirement estimates for pigs weighing between 20 and 90 kg have varied from 6.0 to 10.5 mg of pantothenic acid/kg of diet (Cartron et al., 1953; Pond et al., 1960; Davey and Steven-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins son, 1963; Palm et al., 1968; Meade et al., 1969; RothMaier and Kirchgessner, 1977). Ullrey et al. (1955), Davey and Stevenson (1963), and Teague et al. (1970) reported poor reproductive performance in three experiments when the pantothenic acid level was below 5.9 mg/kg of diet; Bowland and Owen (1952), however, reported normal reproductive performance at this level. Ullrey et al. (1955) and Davey and Stevenson (1963) estimated the pantothenic acid requirement for optimal reproduction at 12.0 to 12.5 mg/kg of diet. Pantothenic acid deficiency signs include slow growth, anorexia, diarrhea, dry skin, rough hair coat, alopecia, reduced immune response, and an abnormal movement of the hind legs called goose stepping (Hughes and Ittner, 1942; Wintrobe et al., 1943b; Luecke et al., 1948, 1950, 1952; Wiese et al., 1951; Stothers et al., 1955; Harmon et al., 1963). Postmortem findings in pigs with pantothenic acid deficiency include edema and necrosis of the intestinal mucosa, increased connective tissue invasion of the submucosa, loss of nerve myelin, and degeneration of dorsal root ganglion cells (Wintrobe et al., 1943b; Follis and Wintrobe, 1946). Riboflavin A component of two coenzymes, flavin mononucleotide (FMN) and flavin adenine dinucleotide (FAD), riboflavin is important in the metabolism of proteins, fats, and carbohydrates. In feedstuffs, most of the riboflavin activity exists as FAD. Estimates of the riboflavin requirement for pigs weighing 2 to 20 kg range from 2.0 to 3.0 mg/kg of synthetic diet (Forbes and Haines, 1952; Miller et al., 1954). Riboflavin requirement estimates range from 1.1 to 2.9 mg/kg for growing pigs fed synthetic diets (Hughes, 1940a; Krider et al., 1949; Mitchell et al., 1950; Terrill et al., 1955), whereas the estimates vary from 1.8 to 3.1 mg/kg of diet when practical diets are fed (Krider et al., 1949; Miller and Ellis, 1951). Seymour et al. (1968) reported no consistent interactions between riboflavin level and environmental temperature for 5- to 17-kg pigs, a finding that contradicted an earlier report by Mitchell et al. (1950). Corn–soybean meal diets are deficient in bioavailable riboflavin. In a study with chicks, Chung and Baker (1990) estimated that the riboflavin in corn–soybean meal diets is 59 percent bioavailable relative to crystalline riboflavin. Riboflavin deficiency has led to anestrus (Esch et al., 1981) and reproductive failure in gilts (Miller et al., 1953; Frank et al., 1984). On the basis of farrowing performance and erythrocyte glutathione reductase activity (FADdependent enzyme), Frank et al. (1984) estimated the available riboflavin requirement for pregnancy to be about 6.5 mg daily. Pettigrew et al. (1996), however, observed that 60 mg of riboflavin/day produced a higher farrowing rate

79

than 10 mg/day when these levels were fed from breeding to day 21 of gestation. Erythrocyte glutathione reductase activity and farrowing performance suggest a lactation requirement of about 16 mg of riboflavin daily (Frank et al., 1988). Signs of riboflavin deficiency in young growing pigs include slow growth, cataracts, stiffness of gait, seborrhea, vomiting, and alopecia (Wintrobe et al., 1944; Miller and Ellis, 1951; Lehrer and Wiese, 1952; Miller et al., 1954). In severe riboflavin deficiency, researchers have observed increased blood neutrophil granulocytes, decreased immune response, discolored liver and kidney tissue, fatty liver, collapsed follicles, degenerating ova, and degenerating myelin of the sciatic and brachial nerves (Wintrobe et al., 1944; Krider et al, 1949; Mitchell et al., 1950; Forbes and Haines, 1952; Lehrer and Wiese, 1952; Miller et al., 1954; Terrill et al., 1955; Harmon et al., 1963). Thiamin Thiamin is essential for carbohydrate and protein metabolism. The coenzyme, thiamin pyrophosphate, is essential for the oxidative decarboxylation of a–keto acids. Thiamin is very heat-labile. Therefore, excess heat or autoclaving can reduce the thiamin content of dietary components, particularly when reducing sugars are present. Miller et al. (1955) estimated a thiamin requirement of 1.5 mg/kg for pigs weighing about 2 kg initially and fed to approximately 10 kg of body weight. Pigs weaned at 3 weeks and fed to about 40 kg of body weight required about 1.0 mg of thiamin/kg of diet (Van Etten et al., 1940; Ellis and Madsen, 1944). The survival time of thiamindeficient pigs was increased by increasing fat levels to 28 percent of the diet (Ellis and Madsen, 1944). This finding indicated that the requirement for thiamin was decreased as the dietary energy from carbohydrate was replaced with higher levels of fat. Weight gain was improved by increasing thiamin levels to 1.1 mg/kg of diet, whereas feed intake was maximized at 0.85 mg/kg of diet for pigs weighing about 30 kg and fed to 90 kg of body weight (Peng and Heitman, 1974). Peng and Heitman (1973) evaluated the thiamin status of growing-finishing pigs by measuring the increase in erythrocyte transketolase activity resulting from thiamin pyrophosphate addition to in vitro preparations. This criterion yielded thiamin requirement estimates up to four times the level required for maximum weight gain. Furthermore, the requirement measured by this criterion increased as environmental temperature increased from 20 to 35°C (Peng and Heitman, 1974). This change was probably related to a reduction in feed intake. There is a lack of information on the thiamin requirement for pregnancy and lactation. Treatment of feed ingredients with sulfur dioxide inactivates thiamin. This process was used in early studies to

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

80

Nutrient Requirements of Swine

produce deficient diets for purposes of determining a pig’s thiamin requirement (Van Etten et al., 1940; Ellis and Madsen, 1944). A number of freshwater fish species contain an antithiamin factor known as thiaminase I (Tanphaichitr and Wood, 1984). Feeding moderate levels of unprocessed freshwater fish preparations to other animals can cause a thiamin deficiency (Green et al., 1941; Krampitz and Woolley, 1944). Thiamin-deficient pigs exhibit loss of appetite; a reduction in weight gain, body temperature, and heart rate; and, occasionally, vomiting. Other effects observed in thiamin deficiency are heart hypertrophy, flabby heart, myocardial degeneration, and sudden death because of heart failure. Animals deficient in thiamin also have elevated plasma pyruvate concentrations (Hughes, 1940b; Van Etten et al., 1940; Follis et al., 1943; Wintrobe et al., 1943a; Ellis and Madsen, 1944; Heinemann et al., 1946; Miller et al., 1955). Most of the cereal grains used in swine diets are rich in thiamin. Hence, grain–oilseed meal diets fed to all classes of swine are considered adequate in this B-vitamin, and it is not generally included as a supplement for swine diets. Vitamin B6 (The Pyridoxines) Vitamin B6 occurs in feedstuffs as pyridoxine, pyridoxal, pyridoxamine, and pyridoxal phosphate. Pyridoxal phosphate is an important cofactor for many amino acid enzyme systems, including transaminases, decarboxylases, dehydratases, synthetases, and racemases. Vitamin B6 plays a crucial role in central nervous system function. It is involved in the decarboxylation of amino acid derivatives for the synthesis of neurotransmitters and neuroinhibitors. Vitamin B6 in corn and soybean meal is about 40 and 60 percent bioavailable for the chick, respectively (Yen et al., 1976). Presumably, it is the same in pigs, although data ¨ are not available. Miller et al. (1957) and Kosters and Kirchgessner (1976a,b) suggested a dietary requirement of 1.0 to 2.0 mg/kg of diet for the pig weighing initially about 2 kg and fed to 10 kg of body weight. Requirement estimates for the 10- to 20-kg pig range from 1.2 to ¨ 1.8 mg of vitamin B6/kg of diet (Sewall et al., 1964; Kosters and Kirchgessner, 1976a,b). Ritchie et al. (1960) reported no treatment differences in reproductive or lactation performance in gilts and sows fed diets containing total pyroxidine levels of either 1.0 or 10.0 mg/kg from the second month of pregnancy through day 35 of lactation. Easter et al. (1983) reported an increase in litter size at birth and at weaning when 1.0 ppm of pyridoxine was added to a corn–soybean meal diet fed to gilts during pregnancy. In another study, the coefficients of glutamic-oxaloacetic transaminase activity in red blood cells of sexually mature gilts fed 0.45 and 2.1 mg of vitamin B6/day were elevated compared with those of gilts fed an excess level of 83 mg of vitamin B6/day. Whole muscle

glutamic-oxaloacetic transaminase activity was reduced in deficient gilts; this reduction suggests that the daily requirement for vitamin B6 may be greater than 2.1 mg (Russell et al., 1985a,b). A deficiency of vitamin B 6 will reduce appetite and growth rate. Advanced deficiency will result in an exudate development around the eyes, convulsions, ataxia, coma, and death. Blood samples from deficient pigs show a reduction in hemoglobin, red blood cells, and lymphocyte counts. Serum iron and gamma globulin are increased. Peripheral myelin and axis cylinder degeneration of the sensory neurons, microcytic hypochromic anemia, and fat infiltration of the liver are characteristic of vitamin B6 deficiency (Hughes and Squibb, 1942; Wintrobe et al., 1942, 1943c; Follis and Wintrobe, 1946; Lehrer et al., 1951; Miller et al., 1957; Harmon et al., 1963). A tryptophan-loading test, in which the conversion of tryptophan to niacin is impaired, can determine vitamin B6 status. This impairment results in elevated xanthurenic acid and kynurenic acid concentrations in the urine (Cartwright et al., 1944). Supplementation of grain–soybean meal diets with vitamin B6 is generally unnecessary, because the level of bioavailable vitamin B6 in feed ingredients will meet the pig’s requirement. Vitamin B12 Vitamin B12, or cyanocobalamin, contains the trace element cobalt in its molecule, which is a unique feature among vitamins. Vitamin B12 as a coenzyme is involved in the de novo synthesis of labile methyl groups derived from formate, glycine, or serine, and their transfer to homocysteine to form methionine. It is also important in the methylation of uracil to form thymine, which is converted to thymidine and used for the synthesis of DNA. Pigs require vitamin B12, but responses to dietary supplementation have been variable. Synthesis of vitamin B12 by microorganisms in the environment and within the intestinal tract as well as the pig’s inclination toward coprophagy may supply sufficient vitamin B12 to satisfy the pig’s requirement (Bauriedel et al., 1954; Hendricks et al., 1964). Ingredients of plant origin are devoid of vitamin B12, but animal and fermentation by-products contain the vitamin. In these ingredients, vitamin B12 exists in a methylated form (methylcobalamin) or a 58-deoxyadenosyl form (adenosyl cobalamin), and both of these compounds are generally bound to protein. Vitamin B12 supplements are produced commercially by microbial fermentation and are usually added to grain–soybean meal diets. Receptor sites for vitamin B12 binding are located in the ileum. Prior to absorption, cobalamin is bound to a glycoprotein, commonly referred to as ‘‘intrinsic factor.’’ Intrinsic factor is derived from the parietal cells of gastric mucosa. Vitamin B12 is stored effectively in the body. Thus tissue storage, primarily in the liver, resulting from excess

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins vitamin B12 ingestion can delay for many months the onset of vitamin B12 deficiency symptoms after a vitamin B12deficient diet is fed. Estimated vitamin B12 requirements for 1.5- to 20-kg pigs fed synthetic milk diets and housed in wire-floored cages range from 15 to 20 mg/kg of dietary dry matter (Anderson and Hogan, 1950b; Nesheim et al., 1950; Frederick and Brisson, 1961), but as high as 50 mg/kg of diet dry matter in one study (Neumann et al., 1950). Pigs weighing about 10 to 45 kg required 8.8 to 11.0 mg of vitamin B12/ kg of diet (Richardson et al., 1951; Catron et al., 1952). The animals in these experiments also were housed in wirefloored cages. Anderson and Hogan (1950a), Frederick and Brisson (1961), and Teague and Grifo (1966) improved the reproductive performance of sows by adding 11 to 1,100 mg of vitamin B12/kg of diet. Teague and Grifo (1966) compared the reproductive performance of sows fed an unsupplemented all-plant diet with that of a diet supplemented with 110 to 1,100 mg/kg of vitamin B12. Until the sows’ third and fourth parities, there was no reduction in the number of pigs farrowed or weaned, or in their weights at birth or weaning. Because of the wide range of levels supplemented and the few experiments, it is difficult to determine the vitamin B12 requirement for reproduction and lactation, but it is estimated at 15 mg/kg of diet. Pigs that are deficient in vitamin B12 display reduced weight gain, loss of appetite, rough skin and hair coat, irritability, hypersensitivity, and hind leg incoordination. Blood samples from deficient pigs indicate normocytic anemia and high neutrophil and low lymphocyte counts (Anderson and Hogan, 1950b; Neumann and Johnson, 1950; Neumann et al., 1950; Cartwright et al., 1951; Richardson et al., 1951; Catron et al., 1952). A deficiency of folic acid and vitamin B12 has led to macrocytic anemia and bone marrow hyperplasia, both of which have several similar characteristics to pernicious anemia in human beings (Johnson et al., 1950; Cartwright et al., 1952). Signs of folacin deficiency generally accompany vitamin B12 deficiency, because vitamin B12 is required for folate metabolism. Lack of either folacin or vitamin B12 prevents the proper transfer of methyl groups in the synthesis of thymidine. Vitamin C (Ascorbic Acid) Vitamin C (ascorbic acid) is a water-soluble antioxidant that is involved in the oxidation of aromatic amino acids, synthesis of norepinephrine and carnitine, and in the reduction of cellular ferritin iron for transport to the body fluids. Ascorbic acid is also essential for hydroxylation of proline and lysine, which are integral constituents of collagen. Collagen is essential for growth of cartilage and bone. Vitamin C enhances the formation of both bone matrix and tooth dentin. In vitamin C deficiency, petechial hemor-

81

rhages occur throughout the body. A dietary source of vitamin C is essential for primates and guinea pigs, but farm animals, including pigs, can synthesize this vitamin from D -glucose and several other related compounds (Braude et al., 1950; Dvorak, 1974; Brown and King, 1977). Strittmatter et al. (1978), Cleveland et al. (1983), and Nakano et al. (1983) have investigated the role of vitamin C in the prevention or alleviation of osteochondrosis in swine. These authors postulated that osteochondrosis might be related to insufficient collagen cross-linking because of reduced hydroxylation of lysine. Dietary supplementation with vitamin C, however, was ineffective in preventing this malady. Under some conditions, pigs may not be able to synthesize vitamin C rapidly enough to meet their requirements. Riker et al. (1967) reported that plasma ascorbic acid concentrations were lower for pigs at an environmental temperature of 29°C than for pigs at 18°C. However, vitamin C supplementation of pigs housed at temperatures of either 19 or 27°C did not improve rate or efficiency of weight gain (Kornegay et al., 1986). Brown et al. (1970) found a significant correlation between energy intake and serum ascorbate levels, and later reported that vitamin C supplementation significantly improved the rate of weight gain of 3-week-old pigs (Brown et al., 1975). There was a greater response to vitamin C at a low-energy intake than at an intermediate- or a high-energy intake. The concentration and total amount of vitamin C in the liver of 1- or 40-dayold pigs was reduced in fasted pigs compared with that in suckling pigs (Dvorak, 1974). There also are reports of improved weight gains in response to supplemental vitamin C in the diet when no deliberate stress had been imposed on pigs. Jewell et al. (1981) reported improved weight gain from vitamin C supplementation in 1-day-old weaned pigs in one trial, but no response to the supplement in a second trial. Using pigs weaned at 3 to 4 weeks of age, Brown et al. (1975), Yen and Pond (1981), and Mahan et al. (1994) reported that weight gains were improved by supplementing the diet with vitamin C. In pigs weighing 24 kg initially, Mahan et al. (1966) observed an improvement in weight gain from parenteral dosing and feed supplementation with vitamin C. In two of three trials, growing pigs (15 to 27 kg) fed to about 90 kg of body weight responded to vitamin C supplementation (Cromwell et al., 1970). Others have noted no improvement in performance from vitamin C supplementation in suckling pigs, pigs weaned at 3 to 4 weeks of age, or growing-finishing pigs (Hutagalung et al., 1969; Leibbrandt, 1977; Strittmatter et al., 1978; Mahan and Saif, 1983; Nakano et al., 1983; Yen and Pond, 1984; Yen et al., 1985; Kornegay et al., 1986). Mahan et al. (1994) observed no beneficial effects from adding ascorbate to corn–soybean meal diets fed to growing-finishing pigs. Chiang et al. (1985) reviewed the effects of supplemental vitamin C for weanling and growing-finishing pigs.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

82

Nutrient Requirements of Swine

Sandholm et al. (1979) reported a rapid cessation of navel bleeding in newborn pigs when 1.0 g of vitamin C/ day was fed to pregnant sows beginning 5 days before expected farrowing. Pigs from sows given supplemental ascorbic acid were significantly heavier at 3 weeks of age than those from control sows. A water-soluble vitamin K administered in the drinking water to several sows in this herd failed to prevent the navel bleeding problem in newborn pigs. In subsequent studies, there was no improvement in pig survival or growth rate when sows were supplemented with 1.0 to 10.0 g of ascorbic acid/day beginning in late pregnancy (Lynch and O’Grady, 1981; Chavez, 1983; Yen and Pond, 1983). Navel bleeding was not considered to be a problem in these latter experiments. Currently, the conditions in which supplemental vitamin C may be beneficial are not well defined. Therefore, no vitamin C requirement estimate is given for pigs.

REFERENCES Adams, C. R., C. E. Richardson, and T. J. Cunha. 1967. Supplemental biotin and vitamin B6 for swine. J. Anim. Sci. 26:903. (Abstr.) Adamstone, P. B., J. D. Krider, M. F. James, and C. A. Blomquist. 1949. Response of swine to vitamin E-deficient rations. Ann. N.Y. Acad. Sci. 52:260–268. Agricultural Research Council. 1981. The Nutrient Requirements of Pigs. Slough, England: Commonwealth Agricultural Bureaux. Ames, S. R. 1979. Biopotencies in rats of several forms of alpha-tocopherol. J. Nutr. 109:2198–2204. Anderson, G. C., and A. C. Hogan. 1950a. Adequacy of synthetic diets for reproduction of swine. Proc. Soc. Exp. Biol. Med. 75:288–290. Anderson, G. C., and A. C. Hogan. 1950b. Requirements of the pig for vitamin B12. J. Nutr. 40:243–250. Anderson, L. E., R. O. Myer, J. H. Brendemuhl, and L. R. McCowell. 1995a. Bioavailability of various vitamin E compounds for finishing swine. J. Anim. Sci. 73:490–495. Anderson, L. E., R. O. Myer, J. H. Brendemuhl, and L. R. McCowell. 1995b. The effect of excessive dietary vitamin A on performance and vitamin E status in swine fed diets varying in dietary vitamin E. J. Anim. Sci. 73:1093–1098. Anderson, M. D., V. C. Speer, J. T. McCall, and V. W. Hays. 1966. Hypervitaminosis A in the young pig. J. Anim. Sci. 25:1123–1127. Anderson, P. A., D. H. Baker, and S. P. Mistry. 1978. Bioassay determination of the biotin content of corn, barley, sorghum and wheat. J. Anim. Sci. 47:654–659. Anderson, P. A., D. H. Baker, P. A. Sherry, and J. E. Corbin. 1979. Choline-methionine interrelationship in feline nutrition. J. Anim. Sci. 49:522–527. Anonymous. 1990. Nomenclature policy: Generic descriptors and trivial names for vitamins and related compounds. J. Nutr. 120:12–20. Baker, D. H. 1991. Vitamin bioavailability. Pp. 399–431 in Bioavailability of Nutrients for Animals: Amino Acids, Minerals, and Vitamins. Ammerman, C. B., D. H. Baker, and A. J. Lewis, eds. San Diego, CA: Academic Press. Barnhart, C. E., D. V. Catron, G. C. Ashton, and L. Y. Quinn. 1957. Effects of dietary pantothenic acid levels on the weanling pig. J. Anim. Sci. 16:396–403.

Bauriedel, W. R., A. B. Hoerlein, J. C. Picken, Jr., and L. A. Underkofler. 1954. Selection of diet for studies of vitamin B12 depletion using unsuckled baby pigs. J. Agric. Food Chem. 2:468–471. Baustad, B., and I. Nafstad. 1972. Hematologic response to vitamin E in piglets. Br. J. Nutr. 28:183–190. Bengtsson, G., J. Hakkarainen, L. Jonsson, J. Lannek, and P. Lindberg. 1978a. Requirement for selenium (as selenite) and vitamin E (as atocopherol) in weaned pigs. I. The effect of varying a-tocopherol levels in a selenium deficient diet on the development of the VESD syndrome. J. Anim. Sci. 46:143–152. Bengtsson, G., J. Hakkarainen, L. Jonsson, J. Lannek, and P. Lindberg. 1978b. Requirement for selenium (as selenite) and vitamin E (as atocopherol) in weaned pigs. II. The effect of varying selenium levels in a vitamin E deficient diet on the development of the VESD syndrome. J. Anim. Sci. 46:153–160. Bethke, R. M., W. Burroughs, O. H. M. Wilder, B. H. Edgington, and W. L. Robison. 1946. The comparative efficiency of vitamin D from irradiated yeast and cod liver oil for growing pigs, with observations on their vitamin D requirements. Ohio Agricultural Experiment Station Bulletin 667:1–29. Wooster, Ohio: Ohio Agricultural Experiment Station. Bieri, J. G., and M. C. McKenna. 1981. Expressing dietary values for fatsoluble vitamins: Changes in concepts and terminology. Am. J. Clin. Nutr. 34:289–295. Boass, A., S. U. Toverud, T. A. McCain, J. W. Pike, and M. R. Haussler. 1977. Elevated serum levels of 1,g-25-dihydroxycholecalciferol in lactating rats. Nature 267:630–632. Bonnette, E. D., E. T. Kornegay, M. D. Lindemann, and C. Hammerberg. 1990. Humoral and cell-mediated immune response and performance of weaned pigs fed four supplemental vitamin E levels and housed at two nursery temperatures. J. Anim. Sci. 68:1337–1345. Bowland, J. P., and B. D. Owen. 1952. Supplemental pantothenic acid in small grain rations for swine. J. Anim. Sci. I 1: 757. (Abstr.) Boyd, R. D., B. D. Moser, E. R. Peo, Jr., A. J. Lewis, and R. K. Johnson. 1982. Effect of tallow and choline chloride addition to the diet of sows on milk composition, milk yield and preweaning pig performance. J. Anim. Sci. 54: 1–7. Braidman, I. P., and D. C. Anderson. 1985. Extra-endocrine functions of vitamin D. Clin. Endocrinol. 23:445–460. Braude, R., A. S. Foot, K. M. Henry, S. K. Kon, S. Y. Thompson, and T. H. Mead. 1941. Vitamin A studies with rats and pigs. Biochem. J. 35:693–707. Braude, R., S. K. Kon, and E. G. White. 1946. Observations on the nicotinic acid requirements of pigs. Biochem. J. 40:843–855. Braude, R., S. K. Kon, and J. W. G. Porter. 1950. Studies in the vitamin C metabolism of the pig. Br. J. Nutr. 4:186–197. Brief, S., and B. P. Chew. 1985. Effects of vitamin A and b-carotene on reproductive performance in gilts. J. Anim. Sci. 60:998–1004. Brooks, C. C., R. M. Nakamura, and A. Y. Miyahara. 1973. Effect of menadione and other factors on sugar-induced heart lesions and hemorrhagic syndrome in the pig. J. Anim. Sci. 37:1344–1350. Brooks, P. H., D. A. Smith, and V. C. R. Irwin. 1977. Biotin supplementation of diets: The incidence of foot lesions and the reproductive performance of sows. Vet. Rec. 101:46–50. Brown, R. G., and G. J. King. 1977. Ascorbic acid synthesis in pigs. Can. J. Anim. Sci. 57:831. (Abstr.) Brown, R. G., V. D. Sharma, and L. G. Young. 1970. Ascorbic acid metabolism in swine. Interrelationships between the level of energy intake and serum ascorbate levels. Can. J. Anim. Sci. 50:605–609. Brown, R. G., J. G. Buchanan-Smith, and V. D. Sharma. 1975. Ascorbic acid metabolism in swine. The effects of frequency of feeding and level of supplementary ascorbic acid on swine fed various energy levels. Can. J. Anim. Sci. 55:353–358.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins Brubacher, G., and O. Wiss. 1972. Vitamin E active compounds, synergists and antagonists. Pp. 255–258 in The Vitamins, Volume V, W. H. Sebrell and R. S. Harris, eds. New York: Academic Press. Bryant, K. L., G. E. Combs, and H. D. Wallace. 1977. Supplemental choline for young and growing-finishing swine. Florida Agricultural Experiment Station Twenty-second Annual Swine Field Day. Research Report AL-1977-1. Gainesville, FL: Florida Agricultural Experiment Station. Bryant, K. L., E. T. Kornegay, J. W. Knight, H. P. Veit, and D. R. Notter. 1985a. Supplemental biotin for swine. III. Influence of supplementation to corn- and wheat-based diets on the incidence and severity of toe lesions, hair and skin characteristics and structural soundness of sows housed in confinement during four parities. J. Anim. Sci. 60:154–162. Bryant, K. L., E. T. Kornegay, J. W. Knight, K. E. Webb, Jr., and D. R. Notter. 1985b. Supplemental biotin for swine. I. Influence on feedlot performance, plasma biotin and toe lesions in developing gilts. J. Anim. Sci. 60:136–144. Bryant, K. L., E. T. Kornegay, J. W. Knight, K. E. Webb, Jr., and D. R. Notter. 1985c. Supplemental biotin for swine. II. Influence of supplementation to corn- and wheat-based diets on reproductive performance and various biochemical criteria of sows during four parities. J. Anim. Sci. 60:145–153. Burroughs, W., B. H. Edginton, W. L. Robinson, and R. M. Bethke. 1950. Niacin deficiency and enteritis in growing pigs. J. Nutr. 41:51–62. Carter, E. G. A., and Carpenter, K. J. 1982. The available niacin values of food for rats and their relation to analytical values. J. Nutr. 112:2091–2103. Cartwright, G. E., M. M. Wintrobe, P. Jones, M. Lauritsen, and S. Humphreys. 1944. Tryptophane derivatives in urine of pyridoxine deficient swine. Bull. Johns Hopkins Hosp. 75:35. Cartwright, G. E., B. Tatting, and M. M. Wintrobe. 1948. Niacin deficiency anemia in swine. Arch. Biochem. 19:109–118. Cartwright, G. E., B. Tatting, H. Ashenbrucker, and M. M. Wintrobe. 1949. Experimental production of nutritional macrocytic anemia in swine. Blood 4:301–323. Cartwright, G. E., J. G. Palmer, B. Tatting, H. Ashenbrucker, and M. M. Wintrobe. 1950. Experimental production of nutritional macrocytic anemia in swine. III. Further studies on pteroylglutamic acid deficiency. J. Lab. Clin. Med. 36:675–693. Cartwright, G. E., B. Tatting, J. Robinson, N. M. Fellows, F. D. Gunn, and M. M. Wintrobe. 1951. Hematologic manifestations of vitamin B12 deficiency in swine. Blood 6:867–891. Cartwright, G. E., B. Tatting, D. Kurth, and M. M. Wintrobe. 1952. Experimental production of nutritional macrocytic anemia in swine. V. Hematologic manifestations of a combined deficiency of vitamin B12 and pteroylglutamic acid. Blood 7:992–1004. Catron, D. V., D. Richardson, L. A. Underkofler, H. M. Maddock, and W. C. Friedland. 1952. Vitamin B12 requirement of weanling pigs. II. Performance on low level of vitamin B12 and requirements for optimum growth. J. Nutr. 47:461–468. Chavez, E. R. 1983. Supplemental value of ascorbic acid during late gestation on piglet survival and early growth. Can. J. Anim. Sci. 63:683–687. Chew, B. P., H. Rasmussen, M. H. Pubols, and R. L. Preston. 1982. Effects of vitamin A and b-carotene on plasma progesterone and uterine protein secretions in gilts. Theriogenology 18:643–654. Chiang, S. H., J. E. Pettigrew, R. L. Moser, S. G. Cornelius, K. P. Miller, and T. R. Heeg. 1985. Supplemental vitamin C in swine diets. Nutr. Rep. Int. 31:573–581. Chung, T. K., and D. H. Baker. 1990. Riboflavin requirement of chicks fed purified amino acid and conventional corn-soybean meal diets. Poult. Sci. 69:1357–1363. Chung, Y. K., D. C. Mahan and A. J. Lepine. 1992. Efficacy of Da–tocopherol and DL-a–tocopheryl acetate for weanling pigs. J. Anim. Sci. 70:2485–2492.

83

Cleveland, E. R., G. L. Newton, B. G. Mullinix, O. M. Hale, and T. M. Frye. 1983. Foot-leg and performance traits of boars fed two levels of ascorbic acid. J. Anim. Sci. 57(Suppl. 1):387 (Abstr.). Cline, J. H., D. C. Mahan, and A. L. Moxon. 1974. Progeny effects of supplemental vitamin E in sow diets. J. Anim. Sci. 39:974. (Abstr.) Coelho, M. B. 1991. Vitamin stability. Feed Management 42 (No. 10): 24. Combs, G. E., T. H. Berry, H. D. Wallace, and R. C. Crum, Jr. 1966. Levels and sources of vitamin D for pigs fed diets containing varying quantities of calcium. J. Anim. Sci. 25:827-830. Copelin, J. L., H. Monegue, and G. E. Combs. 1980. Niacin levels in growing-finishing swine diets. J. Anim. Sci. 51 (Suppl. 1): 190 (Abstr.). Cromwell, G. L., V. W. Hays, and J. R. Overfield. 1970. Effect of dietary ascorbic acid on performance and plasma cholesterol levels of growing swine. J. Anim. Sci. 31:63–66. Cunha, T. J., D. C. Lindley, and M. E. Ensminger. 1946. Biotin deficiency syndrome in pigs fed desiccated egg white. J. Anim. Sci. 5:219–225. Cunha, T. J., L. K. Bustad, W. E. Ham, D. R. Cordy, E. C. McCullock, I. F. Woods, G. H. Corner, and M. A. McGregor. 1947. Folic acid, para-aminobenzoic acid and anti-pernicious anemia liver extract in swine nutrition. J. Nutr. 34:173–187. Cunha, T. J., R. W. Colby, L. K. Bustad, and J. F. Bone. 1948. The need for and interrelationship of folic acid, anti-pernicious anemia liver extract, and biotin in the pig. J. Nutr. 36:215–229. Davey, R. J., and J. W. Stevenson. 1963. Pantothenic acid requirement of swine for reproduction. J. Anim. Sci. 22:9–13. Davis, C. L., and J. R. Gorham. 1954. The pathology of experimental and natural cases of yellow fat disease in swine. Am. J. Vet. Res. 15:55–59. De Ritter, E. 1976. Stability characteristics of vitamins in processed foods. Food Technol. 30:48–54. Dodd, D. C., and P. E. Newling. 1960. Muscle degeneration and liver necrosis in the pig. Report of a natural outbreak. N. Z. Vet. J. 8:95–98. Duel, H. J., Jr., E. R. Meserve, C. H. Johnston, A. Polgar, and L. Zechmeister. 1945. Reinvestigation of the relative provitamin A potency of cryptoxanthin and b-carotene. Arch. Biochem. 7:447–450. Dvorak, M. 1974. Effects of corticotrophin, starvation and glucose on ascorbic acid levels in the blood plasma and liver of piglets. Nutr. Metab. 16:215–222. Easter, R. A., P. A. Anderson, E. J. Michel, and J. R. Corley. 1983. Response of gestating gilts and starter, grower and finisher swine to biotin, pyridoxine, folacin and thiamine additions to a corn-soybean meal diet. Nutr. Rep. Int. 28:945–953. Ellis, N. R., and L. L. Madsen. 1944. The thiamine requirements of pigs as related to the fat content of the diet. J. Nutr. 27:253–292. Ellis, R. P., and M. M. Vorhies. 1976. Effect of supplemental dietary vitamin E on the serologic response of swine to an Esherichia coli bacterin. J. Am. Vet. Med. Assoc. 168:231–232. Emerson, O. H., G. A. Emerson, A. Mohammad, and H. M. Evans. 1937. The chemistry of vitamin E: Tocopherols from various sources. J. Biol. Chem. 122:99–107. Emmert, J. L. 1997. Methyl group metabolism in poultry. Proc. Degussa Tech. Symp. (Indianapolis, IN), pp. 19–40. Emmert, J. L., and D. H. Baker. 1997. A chick bioassay approach for determining the bioavailable choline concentration in normal and overheated soybean meal, canola meal, and peanut meal. J. Nutr. 127:745–752. Emmert, J. L., T. A. Garrow, and D. H. Baker. 1996. Development of an experimental diet for determining bioavailable choline concentration and its application in studies with soybean lecithin. J. Anim. Sci. 74:2738–2744. Esch, M. W., R. A. Easter, and J. M. Bahr. 1981. Effect of riboflavin deficiency on estrous cyclicity in pigs. Biol. Reprod. 25:659–665. Evans, H. M., O. H. Emerson, and G. A. Emerson. 1936. The isolation from wheat germ oil of an alcohol, a-tocopherol, having the properties of vitamin E. J. Biol. Chem. 113:319–332.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

84

Nutrient Requirements of Swine

Ewan, R. C., M. E. Wastell, E. J. Bicknell, and V. C. Speer. 1969. Performance and deficiency symptoms of young pigs fed diets low in vitamin E and selenium. J. Anim. Sci. 29:912–915. Fidge, N. H., F. R. Smith, and D. S. Goodman. 1969. Vitamin A and carotenoids. The enzymatic conversion of b-carotene into retinal in hog intestinal mucosa. Biochem. J. 114:689–694. Firth, J., and B. C. Johnson. 1956. Quantitative relationships of tryptophan and nicotinic acid in the baby pig. J. Nutr. 59:223–234. Follis, R. H., Jr., and M. M. Wintrobe. 1945. A comparison of the effects of pyridoxine and pantothenic acid deficiencies on the nervous tissues of swine. J. Exp. Med. 81:539–551. Follis, R. H., M. H. Miller, M. M. Wintrobe, and H. J. Stein. 1943. Development of myocardial necrosis and absence of nerve degeneration in thiamine deficiency in pigs. Am. J. Pathol. 19:341–357. Forbes, R. M., and W. T. Haines. 1952. The riboflavin requirement of the baby pig. J. Nutr. 47:411–424. Frank, G. R., J. M. Bahr, and R. A. Easter. 1984. Riboflavin requirement of gestating swine. J. Anim. Sci. 59:1567–1572. Frank, G. R., J. M. Bahr, and R. A. Easter. 1988. Riboflavin requirement of lactating swine. J. Anim. Sci. 66:47–52. Frape, D. L., V. C. Speer, V. W. Hays, and D. V. Catron. 1959. The vitamin A requirement of the young pig. J. Nutr. 68:173–187. Frederick, G. L., and G. J. Brisson. 1961. Some observations on the relationship between vitamin B12 and reproduction in swine. Can. J. Anim. Sci. 41:212–219. Frigg, M. 1976. Bio-availability of biotin in cereals. Poult. Sci. 55:2310–2318. Fritschen, R. D., O. D. Grace, and E. R. Peo, Jr. 1971. Bleeding pig disease. Nebr. Swine Rep. EC71 219:22–23. Gannon, N. J., and J. Liebholz. 1989. The effects of folic acid supplementation on the performance of growing pigs. P. 136 in Manipulating Pig Production II, J. L. Barnett and D. P. Hennessy, eds. Werribee, Victoria, Australia: Australasian Pig Science Assn. Goff , J. P., R. L. Horst, and E. T. Littledike. 1984. Effect of sow vitamin D status at parturition on the vitamin D status of neonatal pigs. J. Nutr. 114:163–169. Goodman, D. S. 1979. Vitamin A and retinoids: Recent advances. Fed. Proc. 38:2501–2503. Goodman, D. S. 1980. Vitamin A metabolism. Fed. Proc. 39:2716–2722. Grandhi, R. R., and J. H. Strain. 1980. Effect of biotin supplementation on reproductive performance and foot lesions in swine. Can. J. Anim. Sci. 60:961–969. Grant, C. A. 1961. Morphological and etiological studies of dietetic microangiopathy in pigs (‘‘mulberry heart’’). Acta Vet. Scand. 2(Suppl. 3):1 (Abstr.). Green, J., P. Mamalis, S. Marcinkiewicz, and D. McHale. 1960. Structure of b-tocopherol. Chem. Ind. (London), Jan. 16, p. 73. Green, R. G., W. E. Carlson, and C. A. Evans. 1941. A deficiency disease of foxes produced by feeding fish. J. Nutr. 21:243–256. Greenberg, S. M., A. Chatterjee, C. E. Calbert, H. J. Duel, Jr., and L. Zechmeister. 1950. A comparison of the provitamin A activity of bcarotene and cryptoxanthin in the chick. Arch. Biochem. 25:61–65. Groce, A. W., E. R. Miller, K. K. Keahey, D. E. Ullrey, and D. J. Ellis. 1971. Selenium supplementation of practical diets for growing-finishing swine. J. Anim. Sci. 32:905–911. Groce, A. W., E. R. Miller, D. E. Ullrey, P. K. Ku, K. K. Keahey, and D. J. Ellis. 1973. Selenium requirements in corn-soy diets for growingfinishing swine. J. Anim. Sci. 37:948–956. Guilbert, H. R., R. F. Miller, and E. H. Hughes. 1937. The minimum vitamin A and carotene requirements of cattle, sheep, and swine. J. Nutr. 13:543–564. Hakkarainen, J., P. Lindberg, G. Bengtsson, L. Jonsson, and N. Lannek. 1978. Requirement for selenium (as selenite) and vitamin E (as atocopherol) in weaned pigs. III. The effect on the development of the

VESD syndrome of varying selenium levels in a low-tocopherol diet. J. Anim. Sci. 46: 1001–1008. Hall, D. D., G. L. Cromwell, and T. S. Stahly. 1991. Effects of dietary calcium, phosphorus, calcium:phosphorus ratio and vitamin K on performance, bone strength and blood clotting status of pigs. J. Anim. Sci. 69:646–655. Hall, D. D., G. L. Cromwell, and T. S. Stahly. 1986. The vitamin K requirement of the growing pig. J. Anim. Sci. 63 (Suppl. 1):268 (Abstr.). Halloran, B. P., and H. F. DeLuca. 1979. Vitamin D deficiency and reproduction in rats. Science 204:73–74. Hamilton, C. R., and T. L. Veum. 1984. Response of sows and litters to added dietary biotin in environmentally regulated facilities. J. Anim. Sci. 59:151–157. Hamilton, C. R., and T. L. Veum. 1986. Effect of biotin and(or) lysine additions to corn-soybean meal diets on the performance and nutrient balance of growing pigs. J. Anim. Sci. 62:155–162. Hamilton, C. R., T. L. Veum, D. E. Jewell, and J. A. Siwecki. 1983. The biotin status of weanling pigs fed semipurified diets as evaluated by plasma and hepatic parameters. Int. J. Vitamin Nutr. Res. 53:44–50. Hancock, J. D., E. R. Peo, Jr., A. J. Lewis, J. D. Crenshaw, and B. D. Moser. 1986. Vitamin D toxicity in young pigs. J. Anim. Sci. 63(Suppl. 1):268 (Abstr.). Hanke, H. E., and R. J. Meade. 1971. Biotin and pyridoxine additions to diets for pigs weaned at an early age. 1970–71 Minnesota Swine Research Report H-120. St. Paul: University of Minnesota Press. Hanson, L. E., and I. L. Hathaway. 1948. The fertility of boars fed a vitamin E deficient ration. J. Anim. Sci. 7:528 (Abstr.). Harmon, B. G., E. R. Miller, J. A. Hoefer, D. E. Ullrey, and R. W. Luecke. 1963. Relationship of specific nutrient deficiencies to antibody production in swine. II. Pantothenic acid, pyridoxine or riboflavin. J. Nutr. 79:263–268. Harmon, B. G., D. E. Becker, A. H. Jensen, and D. H. Baker. 1969. Nicotinic acid-tryptophan relationship in the nutrition of the weanling pig. J. Anim. Sci. 28:848–852. Harmon, D. G., D. E. Becker, A. H. Jensen, and D. H. Baker. 1970. Nicotinic acid-tryptophan nutrition and immunologic implications in young swine. J. Anim. Sci. 31:339–342. Harper, A. F., M. D. Lindemann, L. I. Chiba, G. E. Combs, D. L. Handlin, E. T. Kornegay, and L. L. Southern. 1994. An assessment of dietary folic acid levels during gestation and lactation on reproductive and lactational performance of sows: a cooperative study. J. Anim. Sci. 72:2338–2344. Heaney, D. P., J. A. Hoefer, D. E. Ullrey, and E. R. Miller. 1963. Effects of marginal vitamin A intake during gestation in swine. J. Anim. Sci. 22:925–928. Heinemann, W. W., M. E. Ensminger, T. J. Cunha, and E. C. McCulloch. 1946. The relation of the amount of thiamine in the ration of the hog to the thiamine and riboflavin content of the tissue. J. Nutr. 31:107–125. Heinle, R. W., A. D. Welch, and J. A. Pritchard. 1948. Essentiality of both the antipernicious anemia factor of liver and pteroylglutamic acid for hematopoiesis in swine. J. Lab. Clin. Med. 33:1647 (Abstr.). Hendricks, D. G., E. R. Miller, D. E. Ullrey, R. D. Struthers, B. V. Baltzer, J. A. Hoefer, and R. W. Luecke. 1967. b-Carotene vs. retinyl acetate for the baby pig and the effect upon ergocalciferol requirement. J. Nutr. 93:37–43. Hendricks, H. K., H. S. Teague, D. R. Redman, and A. P. Grifo, Jr. 1964. Absorption of vitamin B12 from the colon of the pig. J. Anim. Sci. 23:1036–1038. Hentges, J. F., Jr., R. H. Grummer, P. H. Phillips, and G. Bohstedt. 1952a. The minimum requirement of young pigs for a purified source of carotene. J. Anim. Sci. 11: 266–272. Hitchcock, J. P., E. R. Miller, K. K. Keahey, and D. E. Ullrey. 1978. Effects of arsanilic acid and vitamin E upon utilization of natural or supplemental selenium by swine. J. Anim. Sci. 46:425–435.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins Hjarde, W., A. Neimann-Sorensen, B. Palludan, and P. H. Sorensen. 1961. Investigations concerning vitamin A requirement, utilization and deficiency symptoms in pigs. Acta Agric. Scand. 11:13–53. Hoppe, P. P., F. J. Schoner, and M. Frigg. 1992. Effects of dietary retinol on hepatic retinol storage and on plasma and tissue a-tocopherol in pigs. Internat. J. Vit. Nutr. Res. 62:121–129. Horst, R. L., E. T. Littledike, J. L. Riley, and J. L. Napoli. 1981. Quantitation of vitamin D and its metabolites and their plasma concentrations in five species of animals. Anal. Biochem. 116:189–203. Horst, R. L., J. L. Napoli, and E. T. Littledike. 1982. Discrimination in the metabolism of orally dosed ergocalciferol and cholecalciferol by the pig, rat, and chick. Biochem. J. 204:185–189. Hove, E. L., and H. R. Seibold. 1955. Liver necrosis and altered fat composition in vitamin E-deficient swine. J. Nutr. 56:173–186. Hughes, E. H. 1940a. The minimum requirement of riboflavin for the growing pig. J. Nutr. 20:233–238. Hughes, E. H. 1940b. The minimum requirement of thiamine for the growing pig. J. Nutr. 20:239–241. Hughes, E. H., and N. R. Ittner. 1942. The minimum requirement of pantothenic acid for the growing pig. J. Anim. Sci. 1:116–119. Hughes, E. H., and R. L. Squibb. 1942. Vitamin B6 (pyridoxine) in the nutrition of the pig. J. Anim. Sci. 1:320–325. Hutagalung, R. I., C. H. Chaney, R. D. Wood, and D. G. Waddill. 1968. Effects of nitrates and nitrites in feed on utilization of carotene in swine. J. Anim. Sci. 27:79–82. Hutagalung, R. I., G. L. Cromwell, V. W. Hays, and C. H. Chaney. 1969. Effect of dietary fat, protein, cholesterol and ascorbic acid on performance, serum and tissue cholesterol levels and serum lipid levels of swine. J. Anim. Sci. 29:700–705. Jewell, D. E., J. A. Siwecki, and T. L. Veum. 1981. The effect of dietary vitamin C on performance and tissue vitamin C levels in neonatal pigs. J. Anim. Sci. 53(Suppl. 1):98 (Abstr.). Johnson, B. C., and M. F. James. 1948. Choline deficiency in the baby pig. J. Nutr. 36:339–344. Johnson, D. W., and L. S. Palmer. 1939. Individual and breed variations in pigs on rations devoid of vitamin D. J. Agric. Res. 58:929–940. Johnson, B. C., M. F. James, and J. L. Krider. 1948. Raising newborn pigs to weaning age on a synthetic diet with attempts to produce a pteroylglutamic acid deficiency. J. Anim. Sci. 7:486–493. Johnson, B. C., A. L. Neuman, R. O. Nesheim, M. F. James, J. L. Krider, A. S. Dana, and J. B. Thiersch. 1950. The interrelationship of vitamin B12 and folic acid in the baby pig. J. Lab. Clin. Med. 36:537–546. Kindberg, C. G., and J. W. Suttie. 1989. Effect of various intakes of phylloquinone on signs of vitamin K deficiency and serum liver phylloquinone concentrations in the rat. J. Nutr. 119:175–180. Koch, B. A., D. B. Parrish, and S. Sukhonthasarnpa. 1963. Effects of dietary nitrate and nitrite on growing pigs. J. Anim. Sci. 22:840 (Abstr.). Kodicek, E., R. Braude, S. K. Kon, and K. G. Mitchell. 1956. The effect of alkaline hydrolysis of maize on the availability of its nicotinic acid to the pig. Br. J. Nutr. 10:51–66. Kodicek, E., R. Braude, S. K. Kon, and K. G. Mitchell. 1959. The availability to pigs of nicotinic acid in tortilla baked from maize treated with limewater. Br. J. Nutr. 13:363–384. Kopinski, J. S., J. Leibholz, and L. Bryden. 1989. Biotin studies in pigs. 4. Biotin availability in feedstuffs for pigs and chickens. Br. J. Nutr. 62:773–780. Kormann, A. W., and H. Weiser. 1984. Protective functions of fat-soluble vitamins. Pp. 201–222 in Proc. 37th Nottingham Feed Manufacturer’s Conference, Nottingham, England. London: Butterworth. Kornegay, E. T. 1986. Biotin in swine production: A review. Livestock Prod. Sci. 14:65–89. Kornegay, E. T., and T. N. Meacham. 1973. Evaluation of supplemental choline for reproducing sows housed in total confinement on concrete or in dirt lots. J. Anim. Sci. 37:506–509.

85

¨ Kosters, W. W., and M. Kirchgessner. 1976a. Gewichtenwicklung und Futterverwertung Friiheutwijhnter Ferkel bei untersehiedlicher Vitamin B6-Versorgung. (Growth rate and feed efficiency of early-weaned piglets with varying vitamin B6 supply.) Z. Tierphysiol. Tierernahr. ¨Futtermittelkd. 37:235–246. Kosters, W. W., and M. Kirchgessner. 1976b. Zur Veranderugn des Futterverzeha Friiheutwohnter Ferkel bei Untersehiedlicher Vitamin BeVersorgung. (Change in feed intake of early-weaned piglets in response to different vitamin B6 supply.) Z. Tierphysiol. Tierernahr. Futtermittelkd. 37:247–254. Krampitz, L. O., and D. W. Woolley. 1944. The manner of inactivation of thiamine by fish tissue. J. Biol. Chem. 152:9–17. Krider, J. L., S. W. Terrill, and R. F. VanPoucke. 1949. Response of weanling pigs to various levels of riboflavin. J. Anim. Sci. 8:121–125. Kroening, G. H., and W. G. Pond. 1967. Methionine, choline and threonine interrelationships for growth and lipotropic action in the baby pig and rat. J. Anim. Sci. 26:352–357. Lannek, N., P. Lindberg, G. Nilsson, G. Nordstrom, and K. Orstadius. 1961. Production of vitamin E deficiency and muscular dystrophy in pigs. Res. Vet. Sci. 2:67–72. Lehrer, W. P., Jr., and A. C. Wiese. 1952. Riboflavin deficiency in baby pigs. J. Anim. Sci. 11:244–250. Lehrer, W. P., Jr., A. C. Wiese, P. R. Moore, and M. E. Ensminger. 1951. Pyridoxine deficiency in baby pigs. J. Anim. Sci. 10: 65–72. Lehrer, W. P., Jr., A. C. Wiese, and P. R. Moore. 1952. Biotin deficiency in suckling pigs. J. Nutr. 47:203–212. Leibbrandt, V. D. 1977. Influence of ascorbic acid on suckling pig performance. J. Anim. Sci. 45(Suppl. 1):98 (Abstr.). Lewis, A. J., G. L. Cromwell, and J. E. Pettigrew. 1991. Effects of supplemental biotin during gestation and lactation on reproductive performance of sows: A cooperative study. J. Anim. Sci. 69:207–214. Lindemann, M. D., and E. T. Kornegay. 1986. Folic acid additions to weanling pig diets. J. Anim. Sci. 63(Suppl. 1): 35 (Abstr.). Lindemann, M. D., and E. T. Kornegay. 1989. Folic acid supplementation to diets of gestating-lactating swine over multiple parities. J. Anim. Sci. 67:459–464. Lindemann, M.P., G.L. Cromwell, and H.J. Monegue. 1995. Effects of inadequate and high levels of vitamin fortification on performance of weanling pigs. J. Anim. Sci. 73(Suppl. 1):16 (Abstr.). Lindley, D. C., and T. J. Cunha. 1946. Nutritional significance of inositol and biotin for the pig. J. Nutr. 32:47–59. Livingston, A. L., J. W. Nelson, and G. O. Kohler. 1968. Stability of alpha-tocopherol during alfalfa dehydration and storage. J. Agric. Food Chem. 16:492–495. Long, G. G. 1984. Acute toxicosis in swine associated with excessive intake of vitamin D. J. Am. Vet. Med. Assoc. 184:164–170. Lowry, K. R., O. A. Izquierdo, and D. H. Baker. 1987. Efficacy of betaine relative to choline as a dietary methyl donor. Poult. Sci. 66(Suppl. 1):120 (Abstr.). Luce, W. G., E. R. Peo, Jr., and D. B. Hudman. 1966. Availability of niacin in wheat for swine. J. Nutr. 88:39–44. Luce, W. G., E. R. Peo, Jr., and D. B. Hudman. 1967. Availability of niacin in corn and milo for swine. J. Anim. Sci. 26:76–84. Luce, W. G., D. S. Buchanan, C. V. Maxwell, H. E. Jordan, and R. O. Bates. 1985. Effect of supplemental choline and dichlorvos on reproductive performance of gilts. Nutr. Rep. Int. 32:245–251. Luecke, R. W., W. N. McMillen, F. Thorpe, Jr., and C. Tull. 1947. The relationship of nicotinic acid, tryptophane and protein in the nutrition of the pig. J. Nutr. 33:251–261. Luecke, R. W., W. N. McMillen, F. Thorpe, Jr., and C. Tull. 1948. Further studies on the relationship of nicotinic acid, tryptophane and protein in the nutrition of the pig. J. Nutr. 36:417–424. Luecke, R. W., W. N. McMillen, and F. Thorpe, Jr. 1950. Further studies of pantothenic acid deficiency in weanling pigs. J. Anim. Sci. 9:78–82.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

86

Nutrient Requirements of Swine

Luecke, R. W., J. A. Hoefer, and F. Thorpe, Jr. 1952. The relationship of protein to pantothenic acid and vitamin B12 in the growing pig. J. Anim. Sci. 11:238–243. Luecke, R. W., J. A. Hoefer, and F. Thorpe, Jr. 1953. The supplementary effects of calcium pantothenate and aureomycin in a low-protein ration for weanling pigs. J. Anim. Sci. 12:605–610. Lynch, P. B., and J. F. O’Grady. 1981. Effect of vitamin C (ascorbic acid) supplementation on sows in late pregnancy on piglet mortality. Irish J. Agric. Res. 20:217–219. Lynch, P. B., G. E. Hall, L. D. Hill, E. E. Hatfield, and A. H. Jensen. 1975. Chemically preserved high-moisture corns in diets for growingfinishing swine. J. Anim. Sci. 40:1063–1069. Madsen, A., H. P. Mortensen, W. Hjarde, E. Leebeck, and T. Leth. 1973. Vitamin E in barley treated with propionic acid with special reference to the feeding of bacon pigs. Acta Agric. Scand. Suppl. 19:169–173. Mahan, D. C. 1991. Assessment of the influence of dietary vitamin E on sows and offspring in three parities: Reproductive performance, tissue tocopherol, and effects on progeny. J. Anim. Sci. 69:2904–2917. Mahan, D. C. 1994. Effects of dietary vitamin E on sow reproductive performance over a five-parity period. J. Anim. Sci. 72:2870–2879. Mahan, D. C., and A. L. Moxon. 1978. Effect of increasing the level of inorganic selenium supplementation in the post-weaning diets of swine. J. Anim. Sci. 46:384–390. Mahan, D. C., and L. J. Saif. 1983. Efficacy of vitamin C supplementation for weanling pigs. J. Anim. Sci. 56:631–639. Mahan, D. C., R. A. Pickett, T. W. Perry, T. M. Curtin, W. R. Featherston, and W. M. Beeson. 1966. Influence of various nutritional factors and physical form of feed on esophagogastric ulcers in swine. J. Anim. Sci. 25:1019–1023. Mahan, D. C., A. J. Lepine, and K. Dabrowski. 1994. Efficacy of magnesium-L-ascorbyl-2-phosphate as a vitamin C source for weanling and growing-finishing swine. J. Anim. Sci. 72:2354–2361. Malm, A., W. G. Pond, E. F. Walker, Jr., M. Homan, A. Aydin, and D. Kirtland. 1976. Effect of polyunsaturated fatty acids and vitamin E level of the sow gestation diet on reproductive performance and on level of alpha-tocopherol in colostrum, milk and dam and progeny blood serum. J. Anim. Sci. 42:393–399. Matte, J. J., C. L. Girard, and G. J. Brisson. 1984a. Folic acid and reproductive performance of sows. J. Anim. Sci. 59:1020–1025. Matte, J. J., C. L. Girard, and G. J. Brisson. 1984b. Serum folates during the reproductive cycle of sows. J. Anim. Sci. 59:158–163. Matte, J. J., C. L. Girard, and G. J. Brisson. 1992. The role of folic acid in the nutrition of gestating and lactating primaparous sows. Livestock Prod. Sci. 32:131–148. Meade, R. J. 1971. Biotin and pyridoxine supplementation of diets for growing pigs. 1970–71 Minnesota Swine Research Report H-218. St. Paul, Minn.: University of Minnesota Press. Meade, R. J., L. E. Hanson, H. E. Hanke, K. P. Miller, J. W. Rust, R. S. Grant, and M. E. Tumbleson. 1969. B-vitamin supplementation of conventional diets for growing swine. Minnesota Agricultural Experiment Station Technical Bulletin 263. St. Paul: University of Minnesota Press. Meyer, W. R., D. C. Mahan, and A. L. Moxon. 1981. Value of dietary selenium and vitamin E for weanling swine as measured by performance and tissue selenium and glutathione peroxidase activities. J. Anim. Sci. 52:302–311. Michel, R. L., C. K. Whitehair, and K. K. Keahey. 1969. Dietary hepatic necrosis associated with selenium-vitamin E deficiency in swine. J. Am. Vet. Med. Assoc. 155:50–59. Miller, C. O., and N. R. Ellis. 1951. The riboflavin requirement of growing swine. J. Anim. Sci. 10: 807–812. Miller, C. O., N. R. Ellis, J. W. Stevenson, and R. Davey. 1953. The riboflavin requirement of swine for reproduction. J. Nutr. 51:163–170.

Miller, E. R., R. L. Johnston, J. A. Hoefer, and R. W. Luecke. 1954. The riboflavin requirement of the baby pig. J. Nutr. 52:405–413. Miller, E. R., D. A. Schmidt, J. A. Hoefer, and R. W. Luecke. 1955. The thiamine requirement of the baby pig. J. Nutr. 56:423–430. Miller, E. R., D. A. Schmidt, J. A. Hoefer, and R. W. Luecke. 1957. The pyridoxine requirement of the baby pig. J. Nutr. 62:407–419. Miller, E. R., D. E. Ullrey, C. L. Zutaut, B. V. Baltzer, D. A. Schmidt, B. H. Vincent, J. A. Hoefer, and R. W. Luecke. 1964. Vitamin D2 requirement of the baby pig. J. Nutr. 83:140–148. Miller, E. R., D. E. Ullrey, C. L. Zutaut, J. A. Hoefer, and R. W. Luecke. 1965. Comparisons of casein and soy proteins upon mineral balance and vitamin D2 requirement of the baby pig. J. Nutr. 85:347–353. Misir, R., and R. Blair. 1984. Effect of biotin supplementation of a barleywheat diet on restoration of healthy feet, legs, and skin of biotin deficient sows. J. Res. Vet. Sci. 40:212–218. Misir, R., and R. Blair. 1986. Reproductive performance of gilts and sows as affected by induced biotin deficiency and subsequent dietary biotin supplementation. J. Anim. Physiol. Anim. Nutr. 55:196–208. Mitchell, H. H., B. C. Johnson, T. S. Hamilton, and W. T. Haines. 1950. The riboflavin requirement of the growing pig at two environmental temperatures. J. Nutr. 41:317–337. Molitoris, B. A., and D. H. Baker. 1976. Assessment of the quantity of biologically available choline in soybean meal. J. Anim. Sci. 42:481–489. Muhrer, M. E., R. G. Cooper, C. N. Cornell, and R. D. Thomas. 1970. Diet related hemorrhagic syndrome in swine. J. Anim. Sci. 31:1025 (Abstr.). Myers, G. S., Jr., H. D. Eaton, and J. E. Rousseau, Jr. 1959. Relative value of carotene from alfalfa and vitamin A from a dry carrier fed to lambs and pigs. J. Anim. Sci. 18:288–297. Nafstad, I. 1965. Studies of hematology and bone marrow morphology in vitamin E-deficient pigs. Pathol. Vet. 2:277–287. Nafstad, I. 1973. Some aspects of vitamin E deficiency in pigs. Acta Agric. Scand. 19:31–34. Nafstad, I., and H. J. Nafstad. 1968. An electron microscopic study of blood and bone marrow in vitamin E-deficient pigs. Pathol. Vet. 5:520–537. Nafstad, I., and S. Tollersrud. 1970. The vitamin E-deficiency syndrome in pigs. I. Pathological changes. Acta Vet. Scand. 11:452–480. Nakano, T., F. X. Aherne, and J. R. Thompson. 1983. Effect of dietary supplementation of vitamin C on pig performance and the incidence of osteochondrosis in elbow and stifle joints in young growing swine. Can. J. Anim. Sci. 63:421–428. National Research Council. 1987. Vitamin Tolerance of Animals. Washington, D.C.: National Academy Press. 96 pp. Nelson, E. C., B. A. Dehority, H. S. Teague, V. L. Sanger, and W. D. Pounden. 1962. Effect of vitamin A on some biochemical and physiological changes in swine. J. Nutr. 76:325–332. Nelson, E. C., B. A. Dehority, H. S. Teague, A. P. Grifo, Jr., and V. L. Sanger. 1964. Effect of vitamin A and vitamin A acid on cerebrospinal fluid pressure and blood and liver vitamin A concentration in the pig. J. Nutr. 82:263–268. Nesheim, R. O., and B. C. Johnson. 1950. Effect of a high level of methionine on the dietary choline requirement of the baby pig. J. Nutr. 41:149–152. Nesheim, R. O., J. L. Krider, and B. C. Johnson. 1950. The quantitative crystalline B12 requirement of the baby pig. Arch. Biochem. 27:240–242. Neumann, A. L., and B. C. Johnson. 1950. Crystalline vitamin B12 in the nutrition of the baby pig. J. Nutr. 40:403–414. Neumann, A. L., J. L. Krider, M. R. James, and B. C. Johnson. 1949. The choline requirement of the baby pig. J. Nutr. 38:195–214. Neumann, A. L., J. B. Thiersch, J. L. Krider, M. F. James, and B. C. Johnson. 1950. Requirement of the baby pig for vitamin B12 fed as a concentrate. J. Anim. Sci. 9:83–89. Newcomb, M. D., and G. L. Allee. 1986. Water-soluble vitamins for weanling pigs. J. Anim. Sci. 63 (Suppl. 1):108 (Abstr.).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins Newport, M. J. 1981. A note on the effect of low levels of biotin in milk substitutes for neonatal pigs. Anim. Prod. 33:333–335. Nielsen, H. E., N. J. Hojgaard-Olsen, W. Hjarde, and E. Leerbeck. 1973. Vitamin E content in colostrum and sow’s milk and sow milk yield at two levels of dietary fats. Acta Agric. Scand. Suppl. 19:35–38. Nielsen, H. E., V. Danielsen, M. G. Simesen, C. Gissel-Nielsen, W. Hjarde, T. Leth, and A. Basse. 1979. Selenium and vitamin E deficiency in pigs. I. Influence on growth and reproduction. Acta Vet. Scand. 20:276–288. Nockels, C. F. 1979. Protective effects of supplemental vitamin E against infection. Fed. Proc. 38:2134–2138. Noff, D., and S. Edelstein. 1978. Vitamin D and its hydroxylated metabolites in the rat. Hormone Res. 9:292–300. North Central Region-42 Committee on Swine Nutrition. 1976. Effects of supplemental choline on reproductive performance of sows: A cooperative regional study. J. Anim. Sci. 42:1211–1216. North Central Region-42 Committee on Swine Nutrition. 1980. Effect of supplemental choline on performance of starting, growing and finishing pigs: A cooperative regional study. J. Anim. Sci. 50:99–102. Obel, A. L. 1953. Studies on the morphology and etiology of so-called toxic liver dystrophy (hepatosis dietetica) in swine. Acta Pathol. Microbiol. Scand. Suppl. 94:1–87. Oduho, G. and D. H. Baker. 1993. Quantitative efficacy of niacin sources for chicks: nicotinic acid, nicotinamide, NAD and tryptophan. J. Nutr. 123:2201–2206. Oduho, G. W., T. K. Chung, and D. H. Baker. 1993. Menadione nicotinamide bisulfite is a bioactive source of vitamin K and niacin activity for chicks. J. Nutr. 123:737–743. Oduho, G. W., Y. Han, and D. H. Baker. 1994. Iron deficiency reduces the efficacy of tryptophan as a niacin precursor for chicks. J. Nutr. 124:444–450. Osweiler, G. D. 1970. Porcine hemorrhagic disease. No. AS3531 in Proc. Pork Producers Day. Ames, IA. Iowa State Univ. Press. Palm, B. W., R. J. Meade, and A. L. Melliere. 1968. Pantothenic acid requirement of young swine. J. Anim. Sci. 27:1596–1601. Parrish, D. B., C. E. Aubel, J. S. Hughes, and J. D. Wheat. 1951. Relative value of vitamin A and carotene for supplying the vitamin A requirements of swine during gestation and beginning lactation. J. Anim. Sci. 10:551–559. Partridge, I. G., and M. S. McDonald. 1990. A note on the response of growing pigs to supplemental biotin. Anim. Prod. 50:195–197. Peng, C. L., and H. Heitman, Jr. 1973. Erythrocyte transketolase activity and the percentage stimulation by thiamin pyrophosphate as criteria of thiamin status in the pig. Br. J. Nutr. 30:391–399. Peng, C. L., and H. Heitman, Jr. 1974. The effect of ambient temperature on the thiamin requirement of growing-finishing pigs. Br. J. Nutr. 32:1–9. Pennock, J. F., F. W. Hemming, and J. D. Kerr. 1964. A reassessment of tocopherol chemistry. Biochem. Biophys. Res. Commun. 17:542–548. Penny, R. H. C., R. H. A. Cameron, S. Johnson, P. J. Kenyan, H. A. Smith, A. W. P. Bell, J. P. L. Cole, and J. Taylor. 1981. The influence of biotin supplementation on sow reproductive efficiency. Vet. Rec. 109:80–81. Peo, E. R., Jr., G. F. Wehrebein, B. Moser, P. J. Cunningham, and P. E. Vipperman, Jr. 1970. Biotin supplementation of baby pig diets. J. Anim. Sci. 31:209. (Abstr.) Peplowski, M. A., D. C. Mahan, F. A. Murray, A. L. Moxon, A. H. Cantor, and K. E. Ekstrom. 1980. Effect of dietary and injectable vitamin E and selenium in weanling swine antigenically challenged with sheep red blood cell. J. Anim. Sci. 51:344–351. Pettigrew, J.E., S.M. El-Kandelgy, L.J. Johnston, and G. C. Shurson. 1996. Riboflavin nutrition of sows. J. Anim. Sci. 74:2226–2230. Petzold, E. N., F. W. Quackenbush, and M. McQuistan. 1959. Zeacarotenes, new provitamins from corn. Arch. Biochem. Biophys. 82:117–124.

87

Piatkowski, T. L., D. C. Mahan, A. H. Cantor, A. L. Moxon, J. H. Cline, and A. P. Grifo, Jr. 1979. Selenium and vitamin E in semipurified diets for gravid and non-gravid gilts. J. Anim. Sci. 48:1357–1365. Pike, W. J., J. B. Parker, M. R. Haussler, A. Boass, and S. U. Toverud. 1979. Dynamic changes in circulating 1,25-dihydroxyvitamin D during reproduction in rats. Science 204:1427–1429. Piper, R. C., J. A. Froseth, L. R. McDowell, G. H. Kroening, and I. A. Dyer. 1975. Selenium-vitamin E deficiency in swine fed peas (Pisum sativum). Am. J. Vet. Res. 36:273–281. Pond, W. G., E. Kwong, and J. K. Loosli. 1960. Effect of level of dietary fat, pantothenic acid, and protein on performance of growing-fattening swine. J. Anim. Sci. 19:1115–1122. Poor, C. L., S. D. Miller, G. C. Fahey, R. A. Easter, and J. W. Erdman. 1987. Animal models for carotenoid utilization studies: Evaluation of the chick and the pig. Nutr. Rep. Intl. 36:229–234. Powick, W. C., N. R. Ellis, and C. N. Dale. 1947a. Relationship of corn diets to nicotinic acid deficiency in growing pigs. J. Anim. Sci. 6:395–400. Powick, W. C., N. R. Ellis, L. L. Madsen, and C. N. Dale. 1947b. Nicotinic acid deficiency and nicotinic acid requirement of young pigs on a purified diet. J. Anim. Sci. 6:310–324. Powick, W. C., N. R. Ellis, and C. N. Dale. 1948. Relationship of tryptophane to nicotinic acid in the feeding of growing pigs. J. Anim. Sci. 7:228–232. Quarterman, J., A. C. Dalgarno, A. Adams, B. F. Fell, and R. Boyne. 1964. The distribution of vitamin D between the blood and the liver in the pig, and observations on the pathology of vitamin D toxicity. Br. J. Nutr. 18:65–77. Reid, I. M., R. H. Barnes, W. G. Pond, and L. Krook. 1968. Methionineresponsive liver damage in young pigs fed a diet low in protein and vitamin E. J. Nutr. 95:499–508. Richardson, D., D. V. Catron, L. A. Underkofler, H. M. Maddock, and W. C. Friedland. 1951. Vitamin B12 requirement of male weanling pigs. J. Nutr. 44:371–381. Riker, J. T., III, T. W. Perry, R. A. Pickett, and C. J. Heidenreich. 1967. Influence of controlled temperatures on growth rate and plasma ascorbic acid values in swine. J. Nutr. 92:99–103. Ringarp, N. 1960. Clinical and experimental investigations into a postparturient syndrome with agalactia in sows. Acta Agric. Scand. Suppl. 7. Ritchie, H. D., E. R. Miller, D. E. Ullrey, J. A. Hoefer, and R. W. Luecke. 1960. Supplementation of the swine gestation diet with pyridoxine. J. Nutr. 70:491–496. Roth-Maier, D. A., and M. Kirchgessner. 1977. Utersuchungen zum optimalen Pantothensaurebedarf von Mastschweinen. (Studies on the optimal pantothenic acid requirement of market pigs.) Z. Tierphysiol. Tierernahr. Futtermittelkd. 38:121–131. Russell, L. E., P. J. Bechtel, and R. A. Easter. 1985a. Effect of deficient and excess dietary vitamin B6 on amino transaminase and glycogen phosphorylase activity and pyridoxal phosphate content in two muscles from postpubertal gilts. J. Nutr. 115:1124–1135. Russell, L. E., R. A. Easter, and P. J. Bechtel. 1985b. Evaluation of the erythrocyte aspartate aminotransferase activity coefficient as an indicator of the vitamin B 6 status of postpubertal gilts. J. Nutr. 115:1117–1123. Russett, J. C., J. L. Krider, T. R. Cline, H. L. Thacker, and L. B. Underwood. 1979a. Choline-methionine interactions in young swine. J. Anim. Sci. 49:708–714. Russett, J. C., J. L. Krider, T. R. Cline, and L. B. Underwood. 1979b. Choline requirement of young swine. J. Anim. Sci. 48:1366–1373. Sandholm, M., T. Honkanen-Buzalski, and V. Rasi. 1979. Prevention of navel bleeding in piglets by preparturient administration of ascorbic acid. Vet. Rec. 104:337–338. Schendel, H. E., and B. C. Johnson. 1962. Vitamin K deficiency in the baby pig. J. Nutr. 76: 124–130.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

88

Nutrient Requirements of Swine

Schnoes, H. K., and H. F. DeLuca. 1980. Recent progress in vitamin D metabolism and the chemistry of vitamin D metabolites. Fed. Proc. 39:2723–2729. Seerley, R. W., R. J. Emerick, L. B. Embry, and O. E. Olsen. 1965. Effect of nitrate or nitrite administered continuously in drinking water for swine and sheep. J. Anim. Sci. 24:1014–1019. Seerley, R. W., O. W. Charles, H. C. McCampbell, and S. P. Bertch. 1976. Efficacy of menadione dimethylpyrimidinol bisulfite as a source of vitamin K in swine diets. J. Anim. Sci. 42:599–607. Seerley, R. W., R. A. Snyder, and H. C. McCampbell. 1981. The influence of sow dietary lipids and choline on piglet survival, milk and carcass composition. J. Anim. Sci. 52:542–550. Selke, M. R., C. E. Barnhart, and C. H. Chaney. 1967. Vitamin A requirement of the gestating and lactating sow. J. Anim. Sci. 26:759–763. Sewell, R. F., D. G. Price, and M. C. Thomas. 1962. Pantothenic acid requirement of the pig as influenced by dietary fat. Fed. Proc. 21:468. Sewell, R. F., D. Nugara, R. L. Hill, and W. A. Knapp. 1964. Vitamin B-6 requirement of early-weaned pigs. J. Anim. Sci. 23:694–699. Seymour, E. W., V. C. Speer, and V. W. Hays. 1968. Effect of environmental temperature on the riboflavin requirement of young pigs. J. Anim. Sci. 27:389–393. Sharp, B. A., L. G. Young, and A. A. van Dreumel. 1972a. Dietary induction of mulberry heart disease and hepatosis dietetica in pigs. I. Nutritional aspects. Can. J. Comp. Med. 36:371–376. Sharp, B. A., L. G. Young, and A. A. van Dreumel. 1972b. Effect of supplemental vitamin E and selenium in high-moisture corn diets on the incidence of mulberry heart disease and hepatosis dietetica in pigs. Can. J. Comp. Med. 36:393–397. Sheffy, B. E., N. Drouliscos, J. K. Loosli, and J. P. Willman. 1954. Vitamin A requirements of baby pigs. J. Anim. Sci. 13:999. (Abstr.) Simesen, M. G., P. T. Jensen, A. Basse, G. Gissel-Nielsen, T. Leth, V. Danielson, and H. E. Nielsen. 1982. Clinico-pathologic findings in young pigs fed different levels of selenium, vitamin E and antioxidant. Acta Vet. Scand. 23:295–308. Simmins, P. H., and P. H. Brooks. 1980. The effect of dietary biotin level on the physical characteristics of pig hoof tissue. Anim. Prod. 30:469. (Abstr.) Simmins, P. H., and P. H. Brooks. 1983. Supplementary biotin for sows: Effect on reproductive characteristics. Vet. Rec. 112:425–429. Simmins, P. H., and P. H. Brooks. 1985. Effect of different levels of dietary biotin intake on the hoof horn hardness of the gilt. Anim. Prod. 40:544–545 (Abstr.). Southern, L. L., and D. H. Baker. 1981. Bioavailable pantothenic acid in cereal grains and soybean meal. J. Anim. Sci. 53:403–408. Southern, L. L., D. R. Brown, D. D. Werner, and M. C. Fox. 1986. Excess supplemental choline for swine. J. Anim. Sci. 62:992–996. Stahly, T.S., N.H. Williams, S.G. Swenson, and R.C. Ewan. 1995. Dietary B vitamin needs of high and moderate lean growth pigs fed from 9 to 28 kg body weight. J. Anim. Sci. 73(Suppl. 1):193 (Abstr.). Stern, M. H., C. D. Robeson, L. Weisler, and J. G. Blaxter. 1947. dTocopherol. I. Isolation from soybean oil and properties. J. Am. Chem. Soc. 69:869–874. Stockland, W. L., and L. G. Blaylock. 1974. Choline requirement of pregnant sows and gilts under restricted feeding conditions. J. Anim. Sci. 39:1113–1116. Stothers, S. C., D. A. Schmidt, R. L. Johnston, J. A. Hoefer, and R. W. Luecke. 1955. The pantothenic acid requirement of the baby pig. J. Nutr. 57:47–54. Strittmatter, J. E., D. J. Ellis, M. G. Hogberg, A. L. Trapp, M. J. Parsons, and E. R. Miller. 1978. Effects of vitamin C on swine growth and osteochondrosis. J. Anim. Sci. 47 (Suppl. 1): 16 (Abstr.). Suttie, J. W. 1980. The metabolic role of vitamin K. Fed. Proc. 39:2730–2735.

Suttie, J. W., and C. M. Jackson. 1977. Prothrombin structure, activation and biosynthesis. Physiol. Rev. 57:1–70. Sweeney, P. R., and R. G. Brown. 1972. Ultrastructure changes in muscular dystrophy. 1. Cardiac tissue of piglets deprived of vitamin E and selenium. Am. J. Pathol. 68:479–485. Tanphaichitr, V., and B. Wood. 1984. Thiamin. Present Knowledge in Nutrition. Washington, D.C.: The Nutrition Foundation. Teague, H. S., and A. P. Grifo, Jr. 1966. Vitamin B12 in sow rations. Swine Research. Research Summary 13. Ohio Agricultural Research Development Center, Wooster: Ohio State University Press. Teague, H. S., W. M. Palmer, and A. P. Grifo, Jr. 1970. Pantothenic acid deficiency in the reproducing sow. Ohio Agricultural Research Development Center Animal Science Mimeograph 200. Wooster: Ohio State University Press. Terrill, S. W., C. B. Ammerman, D. E. Walker, R. M. Edwards, H. W. Norton, and D. E. Becker. 1955. Riboflavin studies with pigs. J. Anim. Sci. 14:593–603. Tiege, J., Jr. 1977. The generalized Shwartzman reaction induced by a single injection of endotoxin in pigs fed a vitamin E-deficient commercial diet. Acta Vet. Scand. 18:140–142. Tiege, J., Jr., and H. J. Nafstad. 1978. Ultrastructure of colonic epithelial cells in vitamin E- and selenium-deficient pigs. Acta Vet. Scand. 19:549–560. Tiege, J., Jr., K. Nordstoga, and J. Aurjo. 1977. Influence of diet on experimental swine dysentery. I. Effects of vitamin E- and seleniumdeficient diet supplemented with 6.8% cod liver oil. Acta Vet. Scand. 18:384–396. Tiege, J., Jr., F. Saxegaard, and A. Froslie. 1978. Influence of diet on experimental swine dysentery. 2. Effects of a vitamin E- and seleniumdeficient diet supplemented with 3% cod liver oil, vitamin E or selenium. Acta Vet. Scand. 19:133–146. Tollerz, G. 1973. Vitamin E, selenium and some related compounds and tolerance toward iron in piglets. Acta Agric. Scand. Suppl. 19:184–187. Trapp, A. L., K. K. Keahey, D. L. Whitenack, and C. K. Whitehair. 1970. Vitamin E-selenium deficiency in swine: Differential diagnosis and nature of field problem. J. Am. Med. Assoc. 157:289–300. Tremblay, G. F., J. J. Matte, L. Lemieux, and G. J. Brisson. 1986. Serum folates in gestating swine after folic acid addition to diet. J. Anim. Sci. 63:1173–1178. Tribble, L. R., J. D. Hancock, and D. E. Orr, Jr. 1984. Value of supplemental biotin on reproductive performance of sows in confinement. J. Anim. Sci. 59(Suppl. 1):245 (Abstr.). Ullrey, D. E. 1972. Biological availability of fat-soluble vitamins: Vitamin A and carotene. J. Anim. Sci. 35:648–657. Ullrey, D. E. 1974. The selenium deficiency problem in animal agriculture. Pp. 275–293 in Trace Element Metabolism in Animals No. 2, W. G. Hoekstra, J. W. Suttie, H. E. Ganther, and W. Mertz, eds. Baltimore: University Park Press. Ullrey, D. E. 1981. Vitamin E for swine. J. Anim. Sci. 53:1039–1056. Ullrey, D. E., D. E. Becker, S. W. Terrill, and R. A. Notzold. 1955. Dietary levels of pantothenic acid and reproductive performance of female swine. J. Nutr. 57:401–414. Ullrey, D. E., E. R. Miller, R. D. Struthers, R. E. Peterson, J. A. Hoefer, and H. M. Hall. 1965. Vitamin A activity of fermentation b-carotene for swine. J. Nutr. 85:375–385. Ullrey, D. E., E. R. Miller, D. J. Ellis, D. E. Orr, J. P. Hitchcock, K. K. Keahey, and A. L. Trapp. 1971. Vitamin E (selenium and choline), reproduction and MMA. Pp. 48–51 in Report of Swine Research 148, Michigan State University Agricultural Experiment Station. East Lansing: Michigan State University Press. Van Etten, C., N. R. Ellis, and L. L. Madsen. 1940. Studies on the thiamine requirement of young swine. J. Nutr. 20:607–624. Wahlstrom, R. C., and D. E. Stolte. 1958. The effect of supplemental vitamin D in rations for pigs fed in the absence of direct sunlight. J. Anim. Sci. 17:699–705.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Vitamins Wald, G. 1968. Molecular basis of visual excitement. Science 162:230–239. Washam, R. D., J. E. Sowers, and L. W. DeGoey. 1975. Effect of zincproteinate or biotin in swine starter rations. J. Anim. Sci. 40:179. (Abstr.) Wastell, M. E., D. C. Ewan, M. W. Vorhies, and V. C. Speer. 1972. Vitamin E and selenium for growing and finishing pigs. J. Anim. Sci. 34:969–973. Watkins, K. L., L. L. Southern, and J. E. Miller. 1991. Effect of dietary biotin supplementation on sow reproductive performance and soundness and pig growth and mortality. J. Anim. Sci. 69:201–206. Webb, N. G., R. H. C. Penny, and A. M. Johnston. 1984. The effect of a dietary supplement of biotin on pig hoof horn strength and hardness. Vet. Rec. 114:185–189. Weisman, Y., R. Sapir, A. Harell, and S. Edelstein. 1976. Maternal perinatal interrelationships of vitamin D metabolism in rats. Biochem. Biophys. Acta 428:388–395. Wellenreiter, R. H., D. E. Ullrey, E. R. Miller, and W. T. Magee. 1969. Vitamin A activity of corn carotenes for swine. J. Nutr. 99:129–136. Whitehair, C. K., E. R. Miller, M. Loudenslager, and M. G. Hogberg. 1984. MMA in sows-A vitamin E-selenium deficiency. J. Anim. Sci. 59 (Suppl. 1): 106 (Abstr.). Whitehead, C. C., D. W. Bannister, and J. P. F. D’Mello. 1980. Blood pyruvate carboxylase activity as a criterion of biotin status in young pigs. Res. Vet. Sci. 29:126–128. Whittle, K. J., P. J. Dunphy, and J. F. Pennock. 1966. The isolation and properties of d-tocotrienol from Heuca latex. Biochem. J. 100:138–145. Wiese, A. C., W. P. Lehrer, Jr., P. R. Moore, O. F. Pahnish, and W. V. Hartwell. 1951. Pantothenic acid deficiency in baby pigs. J. Anim. Sci. 10:80–87. Wilkinson, J. E., M. C. Bell, J. A. Bacon, and F. B. Masincupp. 1977a. Effects of supplemental selenium on swine. I. Gestation and lactation. J. Anim. Sci. 44:224–228. Wilkinson, J. E., M. C. Bell, J. A. Bacon, and C. C. Melton. 1977b. Effects of supplemental selenium on swine. II. Growing-finishing. J. Anim. Sci. 44:229–233. Wilson, M. E., J. E. Pettigrew, L. J. Johnston, J. D. Hawton, J. W. Rust, and H. Chester-Jones. 1991a. Provision of additional B-vitamins improves growth rate of weanling pigs. J. Anim. Sci. 69 (Suppl. 1):106 (Abstr.). Wilson, M. E., J. E. Pettigrew, R. D. Walker, H. Chester-Jones, and B. Oeltjenbruns. 1991b. Provision of additional vitamin B12 improved growth rate of weanling pigs. J. Anim. Sci. 69 (Suppl. 1): 359 (Abstr.). Wilson, M. E., J. E. Pettigrew, L. J. Johnston, and H. Chester-Jones. 1992a. Effect of B-vitamin supply upon growth of weanling pigs. J. Anim. Sci. 70 (Suppl. 1):61 (Abstr.). Wilson, M. E., J. E. Pettigrew, G. C. Shurson, L. J. Johnston, H. ChesterJones, and J. D. Jones. 1992b. J. Anim. Sci. 70 (Suppl. 1):233 (Abstr.). Wilson, M. E., M. D. Tokach, R. W. Walker, J. L. Nelssen, R. D. Goodhand, and J. E. Pettigrew. 1993. Influence of high levels of individual B vitamins on starter pig performance. J. Anim. Sci. 71 (Suppl. 1):56 (Abstr.). Wintrobe, M. M., M. H. Miller, R. H. Follis, Jr., H. J. Stein, C. Mushatt, and S. Humphreys. 1942. Sensory neuron degeneration in pigs. IV.

89

Protection afforded by calcium pantothenate and pyridoxine. J. Nutr. 24:345–366. Wintrobe, M. M., R. Alcayaga, S. Humphreys, and R. H. Follis, Jr. 1943a. Electrocardiographic changes associated with thiamine deficiency in pigs. Bull. Johns Hopkins Hosp. 73:169. Wintrobe, M. M., R. H. Follis, Jr., R. Alcayaga, M. Paulson, and S. Humphreys. 1943b. Pantothenic acid deficiency in swine with particular reference to the effects on growth and on the alimentary tract. Bull. Johns Hopkins Hosp. 73:313. Wintrobe, M. M., R. H. Follis, Jr., M. H. Miller, H. J. Stein, R. Alcayaga, S. Humphreys, A. Suksta, and G. E. Cartwright. 1943c. Pyridoxine deficiency in swine with particular reference to anemia, epileptiform convulsions and fatty liver. Bull. Johns Hopkins Hosp. 72:1–25. Wintrobe, M. M., W. Buschke, R. H. Follis, Jr., and S. Humphreys. 1944. Riboflavin deficiency in swine with special reference to the occurrence of cataracts. Bull. Johns Hopkins Hosp. 75:102–110. Wintrobe, M. M., H. J. Stein, R. H. Follis, Jr., and S. Humphreys. 1946. Nicotinic acid and the level of protein intake in the nutrition of the pig. J. Nutr. 30:395–412. Wolke, R. E., S. W. Nielsen, and J. E. Rousseau. 1968. Bone lesions of hypervitaminosis A in the pig. Am. J. Vet. Res. 29:1009–1024. Wood, R. D., C. H. Chaney, D. G. Waddill, and G. W. Garrison. 1967. Effect of adding nitrate or nitrite to drinking water on the utilization of carotene by growing swine. J. Anim. Sci. 26:510–513. Wuryastuti, H., H. D Stowe, R. W. Bull, and E. R. Miller. 1993. Effects of vitamin E and selenium on immune responses of peripheral blood, colostrum, and milk leukocytes of sows. J. Anim. Sci. 71:2464–2472. Yen, J. T., and W. G. Pond. 1981. Effect of dietary vitamin C addition on performance, plasma vitamin C and hematic iron status in weanling pigs. J. Anim. Sci. 53:1292–1296. Yen, J. T., and W. G. Pond. 1983. Response of swine to periparturient vitamin C supplementation. J. Anim. Sci. 56:621–624. Yen, J. T., and W. G. Pond. 1984. Responses of weanling pigs to dietary supplementation with vitamin C or carbadox. J. Anim. Sci. 58:132–137. Yen, J. T., A. H. Jensen, and D. H. Baker. 1976. Assessment of the concentration of biologically available vitamin B6, in corn and soybean meal. J. Anim. Sci. 42:866–870. Yen, J. T., A. H. Jensen, and D. H. Baker. 1977. Assessment of the availability of niacin in corn, soybeans and soybean meal. J. Anim. Sci. 45:269–278. Yen, J. T., R. Lauxen, and T. L. Veum. 1978. Effect of supplemental niacin on finishing pigs fed soybean meal supplemented diets. J. Anim. Sci. 47(Suppl. 1):325. (Abstr.). Yen, J. T., P. K. Ku, W. G. Pond, and E. R. Miller. 1985. Response to dietary supplementation of vitamins C and E in weanling pigs fed low vitamin E-selenium diets. Nutr. Rep. Int. 31:877–885. Young, L. G., A. Lun, J. Pos, R. P. Forshaw, and D. Edmeades. 1975. Vitamin E stability in corn and mixed feed. J. Anim. Sci. 40:495–499. Young, L. G., R. B. Miller, D. E. Edmeades, A. Lun, G. C. Smith, and G. J. King. 1977. Selenium and vitamin E supplementation of highmoisture corn diets for swine reproduction. J. Anim. Sci. 45:1051–1060. Young, L. G., R. B. Miller, D. E. Edmeades, A. Lun, G. C. Smith, and G. J. King. 1978. Influence of method of corn storage and vitamin E and selenium supplementation on pig survival and reproduction. J. Anim. Sci. 47:639–647.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

6 Although water is an important nutrient, there has been surprisingly little research conducted on water requirements of swine since the publication of the previous edition of Nutrient Requirements of Swine (National Research Council, 1988). In the future, greater emphasis will need to be placed on the water requirements of swine, because in some areas of the world, water is becoming an increasingly scarce commodity, whereas in others, excessive water usage has led to problems with slurry disposal (Brooks, 1994).

Water

The water content of a pig varies with its age. Water accounts for as much as 82 percent of the empty body weight (whole body weight less gastrointestinal tract contents) in a 1.5-kg neonatal pig and declines to 53 percent in a 90-kg market hog (Shields et al., 1983). This change with age is principally because the fat content of the pig increases with age and adipose tissue is considerably lower in its water content than is muscle (Georgievskii, 1982).

WATER TURNOVER FUNCTIONS OF WATER

Swine obtain water from three sources: (1) water that is consumed; (2) water that is a component of feedstuffs (typically about 10 to 12 percent of air-dry feed); and (3) water that originates from the breakdown of carbohydrate, fat, and protein (metabolic water). The oxidation of 1 kg of fat, carbohydrate, or protein produces 1,190, 560, or 450 g of water, respectively (National Research Council, 1981). According to Yang et al. (1984), every 1 kg of airdry feed consumed will produce between 0.38 and 0.48 kg (or L) of metabolic water. Water is lost from the body by four routes: (1) the lungs (respiration), (2) the skin (evaporation), (3) the intestines (defecation), and (4) the kidneys (urination). Moisture is continually lost from the respiratory tract during the normal process of breathing. Incoming air is both warmed and moistened as it passes over the lining of the respiratory tract and is expired at approximately 90 percent saturation (Roubicek, 1969). For pigs in a thermoneutral environment (20°C), respiratory water loss has been estimated to be 0.29 and 0.58 L for pigs of 20 and 60 kg body weight (Holmes and Mount, 1967). The degree of loss is affected by both temperature and relative humidity; water loss increases with increased temperature and decreases with increased humidity.

Water fulfills a number of physiological functions necessary for life (Roubicek, 1969). It is a major structural element giving form to the body through cell turgidity, and it plays a crucial role in temperature regulation. The high specific heat of water makes it indispensable for dispersing the surplus heat produced during various metabolic processes. About 580 calories of heat are released when 1 g of water changes from liquid to vapor (Thulin and Brumm, 1991). Water is important in the movement of nutrients to the cells of body tissues and for the removal of waste products from these cells. The high dielectric constant of water gives it the ability to dissolve a wide variety of substances and transport these throughout the body via the circulatory system. In addition, water plays a role in virtually every chemical reaction that takes place in the body. The oxidation of carbohydrates, fats, and proteins all result in the formation of water. The subsequent metabolism of these compounds to yield their energy is achieved through a series of complex reactions that include hydration and hydrolysis. Finally, water is important in the lubrication of joints (i.e., synovial fluid) and in providing protective cushioning for the nervous system (i.e., cerebral-spinal fluid).

90

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Water Sweating and insensible water loss from the skin are not major sources of water loss in swine because the sweat glands are largely dormant. Within the thermoneutral zone, the rate of moisture loss has been estimated to be between 12 and 16 g/m2 (Morrison et al., 1967). Increasing the environmental temperature from 15 to 30°C increased water loss from 7 to 32 g/m2 (Ingram, 1964). However, increased relative humidity had no effect on this loss (Morrison et al., 1967). Significant quantities of water are lost in the feces. The amount of manure a pig produces per day in confinement ranges from 8 to 9 percent of its body weight, with a water content varying from 62 to 79 percent (Brooks and Carpenter, 1993). Water loss through the gut will vary with the nature of the diet. In general, the greater the proportion of undigested material, the greater the water loss (Maynard et al., 1979). Water loss increases with the level of fiber intake (Cooper and Tyler, 1959) and with intake of feeds that have laxative properties. Water loss via the feces is also increased in the case of diarrhea (Thulin and Brumm, 1991). Urination is the major route of water excretion in swine, although the amount of water excreted in the urine is highly variable. The kidneys regulate the volume and composition of body fluids by excreting more or less water, depending on water intake and excretion through other mechanisms. Water excretion is increased when pigs are fed diets that contain greater amounts of minerals and protein. The larger the amount of protein in the diet, the greater the water loss, and thus the greater the water requirement (Wahlstrom et al., 1970). Similarly, increased intake of salt results in increases in water intake and a concomitant increase in urinary excretion (Sinclair, 1939).

WATER REQUIREMENTS Many factors, including environmental ones, govern the water requirements of swine (National Research Council, 1981). The amount of water in a pig’s body at any given age is relatively constant. Therefore, pigs must consume sufficient water on a daily basis to balance the amount of water lost. Any factor known to increase water excretion will increase water requirements. The minimum requirement for water is the amount needed to balance water losses, produce milk, and form new tissue during growth or pregnancy. In determining water requirements, care must be taken to distinguish between requirements and consumption. True water usage by pigs is usually overestimated because wastage is generally not taken into account. Based on waterturnover rates measured using tritiated water, water requirements of pigs under confined and normal dry feeding conditions were estimated to be approximately 120 and

91

80 mL/kg of body weight for growing (30 to 40 kg) and non-lactating adult pigs (157 kg), respectively (Yang et al., 1981). However, because of the difficulty in making these types of measurements, water consumption is typically used to estimate water requirement. Suckling Pigs A common assumption is that suckling pigs do not drink water and can completely satisfy their water requirements by drinking milk, because milk contains 80 percent water. However, suckling pigs, in fact, drink water within 1 or 2 days of birth (Aumaitre, 1964). In addition, because milk is a high-protein, high-mineral food, its consumption can cause increased urinary excretion, which might actually lead to a water deficit (Lloyd et al., 1978). As a consequence, research interest in the water requirements of suckling pigs has increased recently. Fraser et al. (1988) measured water use by 51 suckling litters during the first 4 days after farrowing. The use varied greatly among litters, ranging from 0 to 200 mL/day, with an average daily consumption per pig of 46 mL. This level of intake is considerably higher than that reported in earlier work, in which average daily water intake per pig was closer to 10 mL. Fraser et al. (1993) speculated that the increased consumption levels recorded recently may reflect an increased emphasis on temperature control in farrowing rooms and that the higher temperatures currently used may lead to an increase in moisture loss from the pig. Their data showed almost a fourfold increase in water consumption when suckling pigs were housed in rooms at 28°C than when housed at 20°C. Fraser et al. (1988) suggested that providing a supplemental water supply may help to reduce preweaning mortality. They speculated that undernourished pigs, especially those housed in warm environments, may be prone to dehydration during the first few days after farrowing and that at least some pigs have the developmental maturity to compensate by drinking water. Exposed water surfaces (e.g., bowls or cups) are superior to nipple drinkers for this purpose (Phillips and Fraser, 1990, 1991). After the first week of life, the principal concern regarding the water consumption of suckling pigs is the role it plays in stimulating creep feed consumption. Although the consumption of creep feed by pigs is usually low during the first 3 weeks, subsequent feed intake is less if water is not provided (Friend and Cunningham, 1966). Pig health is a factor that affects water intake. Pigs with diarrhea consumed 15 percent less water than healthy pigs (Baranyiova and Holub, 1993). Weanling Pigs Gill et al. (1986) measured the water intake of weaned pigs from 3 to 6 weeks of age. Daily water intake during

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

92

Nutrient Requirements of Swine

the first, second, and third week after weaning averaged 0.49, 0.89, and 1.46 L per pig. The relationship between feed intake and water consumption was described by Brooks et al. (1984) using the following equation. Water intake (L/day) 4 0.149 ` (3.053 2 Daily dry feed intake in kg)

(6-1)

McLeese et al. (1992) observed two distinct patterns of water intake. During the first phase, lasting about 5 days after weaning, water intake fluctuated independently of apparent physiological need and did not seem to be related to growth, feed intake, or the severity of diarrhea. In the second period, water intake followed a consistent pattern that paralleled growth and feed intake. The authors speculated that during the first few days after weaning, water consumption might be high so that the pigs could obtain a sense of satiety in the absence of feed intake. Brooks et al. (1984) reported a diurnal pattern to water intake for weaned pigs housed under conditions of constant light, with a higher consumption from 0830 to 1700 hours than from 0700 to 0830 hours. Nienaber and Hahn (1984) studied the effects of water flow restriction on the performance of weanling pigs. Their results showed little effect on growth when flow rates were varied between 0.1 and 1.1 L/minute. However, water use was significantly higher with a more rapid flow rate, which was attributed to increased wastage of water. Similarly, water use increased when water nipples were tilted up (at 45°) versus down (at 45°) in position (Carlson and Peo, 1982). Weanling pigs in pens with water nipples placed in the down position gained 6.5 percent faster, were 7 percent more efficient in feed conversion, and used 63 percent less water than pigs in pens with water nipples pointing up. There was no advantage in using drip versus non-drip waterers (Ogunbameru et al., 1991). Growing-Finishing Pigs For growing-finishing pigs, free access to water located near feed dispensers is advisable, and such access is normally provided for dry feeding systems. The rate (grams per hour) of digesta or water emptying from the stomach increases as the water intake increases (Low et al., 1985). This process regulates the dry matter content of the gastric digesta, particularly during the first hour after feeding. Factors such as feed intake, ingredients contained in the diet, ambient temperature and humidity, state of health, and stress level affect water requirements. Water consumption generally has a positive relationship with feed intake and body weight (Evvard, 1929). The minimum requirement for pigs between 20 and 90 kg body weight is approximately 2 kg of water for each kg of feed. The voluntary water intake of growing pigs allowed to consume feed ad

libitum is approximately 2.5 kg of water for each kg of feed while pigs receiving restricted amounts of feed have been reported to consume 3.7 kg of water per kg of feed (Cumby, 1986). The difference between ad libitum and restricted fed pigs might be due to the tendency of pigs to fill themselves with water if their appetite is not satisfied by their feed allowance. Braude et al. (1957) gave 79 pigs unrestricted dry feed up to 3 kg/pig daily and free access to water. From 10 to 22 weeks of age, the water-to-feed ratio averaged 2.56:1. From 16 to 18 weeks of age, the maximum average daily intakes of water and feed were 7.0 and 2.7 kg/pig, respectively. Olsson and Andersson (1985), using nose-operated drinking devices, concluded that water consumption at feeding for growing-finishing pigs has a distinct periodicity, with a peak at the beginning and end of the feeding period. Water consumption between feeding periods peaked 2 hours after the morning feeding and 1 hour after the afternoon feeding. These results support the conclusions of Yang et al. (1984) that growing pigs have a tendency, when feed intake is restricted, to increase the total water ingested, possibly because of a desire for abdominal fill. In general, their results suggest that if feed access was restricted, water for abdominal fill was taken during the afternoon. Barber et al. (1988) studied the effect of water delivery rate and number of drinking nipples on the water use of growing pigs. A high (900 mL/minute) delivery rate increased water use (3.8 L/day) compared with a low (300 mL/minute) delivery rate (1.9 L/day). However, pig performance was not affected. Increasing the number of nipples per pen (eight pigs per pen) from one to two had no effect on either water use or pig performance. Mount et al. (1971) reported little difference in water consumption by growing pigs kept at temperatures of 7, 9, 12, 20, or 22°C, although there was considerable variation among pigs at any one temperature. However, at 30 and 33°C, the intake of water increased considerably. At 30°C and above, Close et al. (1971) observed behavioral responses to increased temperature. Urine and feces were voided over the whole pen area, and water was spilled from the water bowl presumably in an attempt to cool the pig’s body surface. The temperature of the water itself will affect intake because additional energy is required to warm liquids consumed at temperatures below that of the body. In an Australian study, pigs were reared from 45 to 90 kg body weight in either a cool room where the temperature was maintained at a constant 22°C or in a hot room where the temperature alternated from 35 to 24°C every 12 hours (Vajrabukka et al., 1981). Pigs kept in the cool room drank 3.3 L daily when the water was cooled to 11°C, compared with almost 4.0 L when the water was warmed to 30°C.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Water In contrast, pigs kept in the hot room drank 10.5 L when the water was supplied at 11°C, but only 6.6 L when it was supplied at 30°C. Hagsten and Perry (1976) reported reductions in water consumption and daily weight gain of 20 and 38 percent, respectively, when growing pigs were fed a diet containing less than 0.20 percent total salt (NaCl) or salt equivalent. Use of antibiotics may also affect water consumption; some researchers report an increase in consumption, whereas others have reported a decrease. It has been hypothesized that the effect of antibiotics on water demand will depend on the relative extent to which water loss is reduced by the control of diarrhea and water demand is increased to enable renal clearance of the antibiotic or its residues (Brooks and Carpenter, 1993). Bowland and Standish (1966) found that withholding access to water for 24 hours before slaughter restricted feed intake and resulted in body weight loss and apparent carcass shrinkage of 5.5 percent and 1.9 kg, respectively. In wet feeding systems, water-to-feed ratios ranging from 1.5:1 to 3.0:1 seemed to have little effect on the performance or carcass quality of growing-finishing swine (Barber et al., 1963; Holme and Robinson, 1965). However, pigs fed with wet feeding systems should be given access to an additional source of fresh water to ensure adequate water intake in case of sudden changes in barn temperature or unexpected alterations in feed composition (e.g., high salt or protein levels). Gestating Sows The water intake of pregnant gilts increases in proportion to dry matter intake (Friend, 1971). For unbred gilts, feed and water intake diminished during estrus (Friend, 1973; Friend and Wolynetz, 1981). Nonpregnant gilts consumed 11.5 L of water daily, whereas gilts in advanced pregnancy consumed 20 L (Bauer, 1982). These quantities are similar to the values of 13.5 L (Riley, 1978) and 10.0 L (Lightfoot and Armsby, 1984). The practice of feed or water deprivation before or after weaning as a means of reducing the weaning-to-breeding interval in sows is not well supported by research evidence (Knabe et al., 1986). According to Madec (1984), urinary disorders are quite common in sows, and low water intake is strongly implicated. Pregnant sows given restricted levels of feed intake may show a desire to compensate for inadequate gut fill by an enhanced water intake. Increasing the fiber content of gestation diets is likely to increase the required ratio of water-to-feed. Lactating Sows Lactating sows need considerable amounts of water, not only to replace the 8 to 16 kg of daily milk secreted but also to void large amounts of metabolic end products in

93

the urine. Daily water consumption for lactating sows was shown to vary from 12 to 40 L/day, with a mean of 18 L/ day (Lightfoot, 1978). These quantities are similar to other recorded values for the daily water intake of lactating sows of 20 L (Bauer, 1982), 25.1 L (Riley, 1978), and 17.7 L (Lightfoot and Armsby, 1984). Phillips et al. (1990) observed no difference in water consumption between sows housed in crates with high (2 L/minute) versus low (0.6 L/minute) flow rates of nipple drinkers. Similarly, the height of the nipple drinkers above the floor (600 mm versus 300 mm) did not affect water consumption patterns. Boars There are few data on the water requirements of boars, but free access to water is advisable. Straub et al. (1976) observed water intakes in boars (70 to 110 kg) of up to 15 L/day at 25°C compared with about 10 L/day at 15°C.

WATER QUALITY Elements and substances can occur in water at levels that are harmful to pigs (National Research Council, 1974). Water may contain a variety of microorganisms, including both bacteria and viruses. Of the former, Salmonella, Leptospira, and Escherichia coli are the most commonly encountered (Fraser et al., 1993). Water can also carry pathogenic protozoa as well as eggs or cysts of intestinal worms. Whether the presence of these microorganisms will be detrimental is largely dependent on the specific types found and their concentration. The Bureau of National Affairs (1973) proposed that water used for livestock should not contain more than 5,000 coliforms/ 100 mL. However, this recommendation can be considered as only a guide because some pathogens may be harmful below this level, whereas other, more benign microorganisms can be tolerated at much higher levels. Bacterial contamination is usually more common in surface waters than in underground supplies such as deep wells and artesian water. Total dissolved solids (TDS) is a measure of the total inorganic matter dissolved in a sample of water. Calcium, magnesium, and sodium in the bicarbonate, chloride, or sulfate form are the most common salts found in water with a high TDS (Thulin and Brumm, 1991). Water containing .6,000 ppm TDS may cause temporary diarrhea and increased daily water intake, although health and performance are not usually affected. Paterson et al. (1979) offered water containing 5,060 ppm TDS to gilts and sows and reported no significant effects on reproduction. The addition of up to 6,000 ppm TDS to water offered to weaned pigs resulted in no effect on growth or feed effi-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

94

Nutrient Requirements of Swine

ciency. However, increases in water intake were reported along with temporary mild diarrhea and less firm feces for pigs offered the higher TDS levels in their water (Anderson and Stothers, 1978; Paterson et al., 1979). Total dissolved solids is an inexact measure of water quality. As a general rule, water containing ,1,000 ppm TDS should be safe, whereas water containing .7,000 ppm TDS may present a health risk for pregnant or lactating sows or for stressed pigs and should not be offered to swine for consumption (National Research Council, 1974). Between 1,000 and 7,000 ppm is a gray area, with some producers reporting economic loss at levels well below 7,000 ppm, whereas others experience transient or minor inconvenience at worst. Since so many different elements can contribute to a high TDS, further chemical analysis should be conducted on such water to determine whether the soluble minerals present represent a health risk. However, the values in Table 6-1 can be used as a guide. The pH of water has little direct relevance to water quality, because almost all samples fall within the acceptable range of 6.5 to 8.5 (Fraser et al., 1993). However, alterations in pH can have a major impact on chemical reactions involved in the treatment of water. High water pH impairs the efficiency of chlorination, and low water pH may cause precipitation of some antibacterial agents delivered via the water system. Sulfonamides particularly pose a risk (Russell, 1985) and could lead to potential problems with carcass sulfa residues, because precipitated medication in the water lines may leach back into the water after medication has been terminated. Water hardness is caused by multivalent metal cations, principally calcium and magnesium. Water is considered soft if hardness is ,60 ppm, hard between 120 and 180 ppm, and very hard .180 ppm (Durfor and Becker, 1964). Even very hard water rarely causes problems for swine (National Research Council, 1980), although it does result in the accumulation of scale in water delivery systems. If this impairs water availability, problems can arise. In one

TABLE 6-1 Evaluation of Water Quality for Pigs Based on Total Dissolved Solids Total Dissolved Solids (ppm)

Rating

Comment

,1,000 1,000 to 2,999

Safe Satisfactory

3,000 to 4,999

Satisfactory

5,000 to 6,999

Reasonable

.7,000

Unfit

No risk to pigs. Mild diarrhea in pigs not adapted to it. May cause temporary refusal of water. Higher levels for breeding stock should be avoided. Risky for breeding stock and pigs exposed to heat stress.

SOURCE:

Adapted from National Research Council, 1974

survey, excessively hard water from one region in Quebec, Canada was found to supply as much as 29 percent of a gestating sow’s daily requirement for calcium (Filpot and Ouellet, 1988). Sulfates are the primary cause of water quality problems in well water in many regions of North America. A recent survey conducted on the Canadian prairies indicated that 25 percent of wells contained excessive (.1,000 ppm) quantities of sulfates (McLeese et al., 1991). Sulfates are not well tolerated in the gut of the pig, resulting in diarrhea and reduced performance when levels are .7,000 ppm (Anderson et al., 1994). However, lower levels (2,650 ppm) have no detrimental effect on pig performance (Maenz et al., 1994). It would seem that pigs can adapt to elevated sulfate levels within a few weeks of exposure. This explains why weanling pigs are most susceptible to sulfates because they consume little water before weaning and, as a consequence, are not adapted. In addition, water odor is not necessarily an indication of poor quality water. Despite a distinct ‘‘rotten egg’’ smell, water containing 1,900 ppm sulfates did not affect pig performance (DeWit et al., 1987). Heavy applications of nitrogenous fertilizers to land and contamination of runoff water by animal wastes can raise nitrate concentrations in water supplies to exceedingly high levels. Nitrites impair the oxygen-carrying capacity of the blood by reducing hemoglobin to methemoglobin. Winks et al. (1950) found that conversion of nitrate to nitrite in the water was necessary for toxicity to occur. They reported mortality in swine with access to well water containing 290 to 490 ppm of nitrate nitrogen. However, Seerley et al. (1965) considered it unlikely that sufficient nitrite would be formed and consumed in water alone to cause toxicity in swine unless the initial level of nitrate exceeds 300 ppm of nitrate nitrogen. Nitrite nitrogen levels greater than 10 ppm are cause for concern (Task Force on Water Quality Guidelines, 1987). Nitrates and nitrites in water also may impair the use of vitamin A by the pig (Wood et al., 1967). Additional ions may be occasionally found in water samples. Safety guidelines are provided in Table 6-2. In situations where poor quality water exists, it is essential to determine its impact on animal performance. Often, producers are overly concerned about the diarrhea in situations where animal performance is not impaired. However, when poor water quality reduces performance, there are a number of things that can be done to alleviate the problem. Chlorination disinfects and destroys disease-causing microorganisms. Protozoa and enteroviruses are much more resistant to chlorination than are bacteria (Fraser et al., 1993). The effectiveness of disinfection and the quantity of chlorine required in the water depends on the quantity of nitrites, iron, hydrogen sulfide, ammonia, and organic matter in the water. The presence of organic matter in the water converts the free chlorine to chloramines, which have less disinfecting action. Sodium hypochlorite or laundry

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Water TABLE 6-2

Water Quality Guidelines for Livestock

Item

Recommended Maximum (ppm) TFWQGa

NRCb

Major ions Calcium Nitrate-N ` Nitrite-N Nitrite-N Sulfate

1,000 100 10 1,000

1 440 33 1

Heavy metals and trace ions Aluminum 5.0 Arsenic 0.5 Beryllium 0.1 Boron 5.0 Cadmium 0.02 Chromium 1.0 Cobalt 1.0 Copper 5.0 Fluoride 2.0 Lead 0.1 Mercury 0.003 Molybdenum 0.5 Nickel 1.0 Selenium 0.05 Uranium 0.2 Vanadium 0.1 Zinc 50.0 a

1 0.2 1 1 0.05 1.0 1.0 0.5 2.0 0.1 0.01 1 1.0 1 1 0.1 25.0

Task Force on Water Quality Guidelines, 1987 National Research Council, 1974

b

bleach (5.25 percent chlorine solution) is commonly used for chlorination. The higher the pH, the more chlorine that is needed to achieve the same degree of disinfection. Some changes in the diet may be warranted in response to problems of water quality. A reduction in the salt (NaCl) level in the diet is common on farms that use water containing a high mineral (TDS) load. Some salt can usually be removed without causing a problem because most diets contain a reasonable safety margin. However, care must be taken to ensure that adequate chloride levels are maintained in the diet because chloride is not usually found in high concentration in poor-quality water. Hard water may be improved with a water softener. The most common type is an ion-exchange unit in which sodium replaces calcium and magnesium in the water. This reduces the hardness of the water but has no effect on the overall mineral load (TDS) because the water then has a higher sodium content. Reverse osmosis units are available to remove sulfates, but both the capital and operating costs of the equipment are prohibitive for a livestock operation.

REFERENCES Anderson, D. M., and S. C. Stothers. 1978. Effects of saline water high in sulfates, chlorides and nitrates on the performance of young weanling pigs. J. Anim. Sci. 47:900–907. Anderson, J. S., D. M. Anderson, and J. M. Murphy. 1994. The effect of water quality on nutrient availability for grower/finisher pigs. Can. J. Anim. Sci. 74: 141–148.

95

´ Aumaitre, A. 1964. Le besoin en eau du porcelet: Etude de la consommation d’eau avant le sevrage. (Water requirements of suckling piglets). Ann. Zootechnol. 13:183–198. Baranyiova, E., and A. Holub. 1993. Effect of diarrhoea on water consumption of piglets weaned on the first day after birth. Acta Vet. Brno 62: 27–32. Barber, R. S., R. Braude, and K. G. Mitchell. 1963. Further studies on the water requirements of the growing pig. Anim. Prod. 5:277–282. Barber, J., P. H. Brooks, and J. L. Carpenter. 1988. The effect of water delivery rate and drinker number on the water use of growing pigs. Anim. Prod. 46:521 (Abstr.). ¨ ¨ Bauer, W. 1982. Der Trankwasserverbrauch guster, hochtragender und laktierender Jungsauen. (Consumption of drinking water by nonpregnant, pregnant and lactating gilts). Arch. Exp. Vet. Med. 36:823–827. Bowland, J. P., and J. F. Standish. 1966. Influence of fasting, water deprivation and stress on carcass shrink of pigs and rats. J. Anim. Sci. 25:377–380. Braude, R., P. M. Clarke, K. G. Mitchell, A. S. Cray, A. Franke, and P. H. Sedgwick. 1957. Unrestricted whey for fattening pigs. J. Agric. Sci. (Camb.) 49:347–356. Brooks, P. H. 1994. Water: Forgotten nutrient and novel delivery system. Pp. 211–234 in Biotechnology in the Feed Industry, P. Lyons and K.A. Jacques, eds. Proceedings of Alltech’s Tenth Annual Symposium. Loughborough, U.K.: Nottingham University Press. Brooks, P. H., and J. L. Carpenter. 1993. The water requirement of growing/finishing pigs: Theoretical and practical considerations. Pp. 179–200 in Recent Developments in Pig Nutrition 2, D.J. Coles, W. Haresign and P.C. Garnsworthy, eds. Loughborough, U.K.: Nottingham University Press. Brooks, P. H., S. J. Russel, and J. L. Carpenter. 1984. Water intake of weaned piglets from three to seven weeks old. Vet. Rec. 115:513–515. Bureau of National Affairs. 1973. EPA drafts water quality criteria as required under federal order law. Environment Reporter 4: 663–670. Carlson, R. L., and E. R. Peo. 1982. Nipple waterer position: Up or Down? Nebraska Swine Report, Lincoln, NE, pp. 8–9. Close, W. H., L. E. Mount, and I. B. Start. 1971. The influence of environmental temperature and plane of nutrition on heat losses from groups of growing pigs. Anim. Prod. 13:285–302. Cooper, P. H., and C. Tyler. 1959. Some effects of bran and cellulose on the water relationships in the digesta and faeces of pigs. Part 1. The effect of including bran and two forms of cellulose in otherwise normal rations. J. Agric Sci. (Cambridge) 52: 332–347. Cumby, T. R. 1986. Design requirements of liquid feeding systems for pigs: A review. J. Agric. Eng. Res. 34: 153–172. DeWit, P., L. G. Young, R. Wenzell, R. Friendship, and D. Peer. 1987. Water quality and pig performance. Can. J. Anim. Sci. 67: 1196 (Abstr.). Durfor, C. M., and E. Becker. 1964. USGS Water-Supply Paper 1812. Washington, DC: U.S. Government Printing Office. Evvard, J. M. 1929. A new feeding method and standards for fattening young swine. Iowa Agricultural Experiment Station Research Bulletin 118. Ames: Iowa State University Press. Filpot, P. M., and G. Ouellet. 1988. Mineral and nitrate content of swine drinking-water in four Quebec regions. Can. J. Anim. Sci. 68: 997–1000. Fraser, D., P. A. Phillips, B. K. Thompson, and W. B. Peeters Weem. 1988. Use of water by piglets in the first days after birth. Can. J. Anim. Sci. 68: 603–610. Fraser, D., J. F. Patience, P. A. Phillips, and J. M. McLeese. 1993. Water for piglets and lactating sows: Quantity, quality and quandraries. Pp. 200–224 in Recent Developments in Pig Nutrition 2, D. J. Coles, W. Haresign, and P. C. Garnsworthy, eds. Loughborough, U.K.: Nottingham University Press. Friend, D. W. 1971. Self-selection of feeds and water by swine during pregnancy and lactation. J. Anim. Sci. 32:658–666.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

96

Nutrient Requirements of Swine

Friend, D. W. 1973. Self-selection of feeds and water by unbred gilts. J. Anim. Sci. 37:1137–1141. Friend, D. W., and H. M. Cunningham. 1966. The effect of water consumption on the growth, feed intake, and carcass composition of suckling piglets. Can. J. Anim. Sci. 46:203–209. Friend, D. W., and M. S. Wolynetz. 1981. Self-selection of salt by gilts during pregnancy and lactation. Can. J. Anim. Sci. 61:429–438. Georgievskii, V. I. 1982. Water metabolism and the animal’s water requirements. Pp. 79–89 in Mineral Nutrition of Animals, V. I. Georgievskii, B. N. Annenkov, and V. I. Samokhin, eds. London: Butterworths. Gill, B. P., P. H. Brooks, and J. L. Carpenter. 1986. The water intake of weaned pigs from 3 to 6 weeks of age. Anim. Prod. 42: 470 (Abstr.). Hagsten, I., and T. W. Perry. 1976. Evaluation of dietary salt levels for swine. 1. Effect on gain, water consumption and efficiency of feed conversion. J. Anim. Sci. 42:1187–1190. Holme, D. W., and K. L. Robinson. 1965. A study of water allowances for the bacon pig. Anim. Prod. 7:377–384. Holmes, C. W., and L. E. Mount. 1967. Heat loss from groups of growing pigs under various conditions of environmental temperature and air movement. Anim. Prod. 9: 435–452. Ingram, D. L. 1964. The effect of environmental temperature on heat loss and thermal insulation in the young pig. Res. Vet. Sci. 5: 357–364. Knabe, D. A., T. J. Prince, and D. E. Orr, Jr. 1986. Effect of feed and (or) water deprivation prior to weaning on reproductive performance of sows: A cooperative study. J. Anim. Sci. 62:1–8. Lightfoot, A. L. 1978. Water consumption of lactating sows. Anim. Prod. 26: 386 (Abstr.). Lightfoot, A. L., and A. W. Armsby. 1984. Water consumption and slurry production of dry and lactating sows. Anim. Prod. 38:541. (Abstr.) Lloyd, L. E., B. E. McDonald, and E. W. Crampton. 1978. Water and its metabolism. Pp. 22–34 in Fundamentals of Nutrition, 2nd Edition. San Francisco: W. H. Freeman and Co. Low, A. G., R. T. Pittman, and R. J. Elliott. 1985. Gastric emptying of barley-soya-bean diets in the pig: Effects of feeding level, supplementary maize oil, sucrose or cellulose, and water intake. Br. J. Nutr. 54:437–447. Madec, F. 1984. Urinary disorders in intensive pig herds. Pig News Info. 5:89–93. Maenz, D. D., J. F. Patience, and M. S. Wolynetz. 1994. The influence of the mineral level in drinking water and thermal environment on the performance and intestinal fluid flux of newly-weaned pigs. J. Anim. Sci. 72: 300–308. Maynard, L. A., J. K. Loosli, H.F. Hintz, and R. G. Warner. 1979. Animal Nutrition. 7th ed. New York: McGraw-Hill. McLeese, J. M., J. F. Patience, M. S. Wolynetz, and G. I. Christison. 1991. Evaluation of the quality of ground water supplies used on Saskatchewan swine farms. Can. J. Anim. Sci. 71: 191–203. McLeese, J. M., M. L. Tremblay, J. F. Patience, and G. I. Christison. 1992. Water intake patterns in the weanling pig: Effect of water quality, antibiotics and probiotics. Anim. Prod. 54: 135–142. Morrison, S. R., T. E. Bond, and H. Heitman. 1967. Skin and lung moisture loss from swine. Trans. Amer. Soc. Agric. Eng. 10: 691–697. Mount, L. E., C. W. Holmes, W. H. Close, S. R. Morrison, and I. B. Start. 1971. A note on the consumption of water by the growing pig at several environmental temperatures and levels of feeding. Anim. Prod. 13:561–563. National Research Council. 1974. Nutrient and Toxic Substances in Water for Livestock and Poultry. Washington, DC: National Academy Press. 93 pp. National Research Council. 1980. Mineral Tolerance of Domestic Animals. Washington, D.C.: National Academy Press.

National Research Council. 1981. Water-environment interactions. Pp. 39–50 in Effect of Environment on Nutrient Requirements of Domestic Animals. Washington, DC: National Academy Press. National Research Council. 1988. Nutrient Requirements of Swine, 9th ed. Washington, DC: National Academy Press. 93 pp. Nienaber, J. A., and G. L. Hahn. 1984. Effects of water flow restriction and environmental factors on performance of nursery-age pigs. J. Anim. Sci. 59:1423–1429. Ogunbameru, B. O., E. T. Kornegay, and C. M. Wood. 1991. A comparison of drip and non-drip nipple waters used by weanling pigs. Can. J. Anim. Sci. 71: 581–583. Olsson, O., and T. Andersson. 1985. Biometric considerations when designing value drinking systems for growing-finishing pigs. Acta Agric. Scand. 35: 55–66. Paterson, D. W., R. C. Wahlstrom, G. W. Libal, and O. E. Olson. 1979. Effects of sulfate in water on swine reproduction and young pig performance. J. Anim. Sci. 49:664–667. Phillips, P. A., and D. Fraser. 1990. Water bowl size for newborn pigs. Appl. Eng. Agric. 6: 79–81. Phillips, P. A., and D. Fraser. 1991. Discovery of selected water dispensers by newborn pigs. Can. J. Anim. Sci. 71: 233–236. Phillips, P. A., D. Fraser, and B. K. Thompson. 1990. The influence of water nipple flow rate and position and room temperature on sow water intake and spillage. Appl. Eng. Agric. 6: 75–78. Riley, J. E. 1978. Drinking ‘‘straws’’: A method of watering housed sows during pregnancy and lactation. Anim. Prod. 26:386 (Abstr.). Russell, I. D. 1985. Some fundamentals of water medications. Poult. Digest 44: 422–423. Seerley, R. W., R. J. Emerick, L. B. Emery, and O. E. Olson. 1965. Effect of nitrate or nitrite administered continuously in drinking water for swine and sheep. J. Anim. Sci. 24:1014–1019. Shields, R. G., Jr., D. C. Mahan, and P. L. Graham. 1983. Changes in swine body composition from birth to 145 kg. J. Anim. Sci. 57:43–54. Sinclair, R. D. 1939. The salt requirements of growing pigs. Sci. Agri. 20: 109–119. Straub, G., J. H. Weniger, E. S. Tawfik, E. S., and D. Steinhauf. 1976. The effects of high environmental temperatures on fattening performance and growth of boars. Livest. Prod. Sci. 3: 65–74. Task Force on Water Quality Guidelines. 1987. Livestock watering. Pp. 4-23–4-37 in Canadian Water Quality Guidelines. Inland Waters Directorate, Ottawa, Ontario. Thulin, A. J., and M. C. Brumm. 1991. Water: The forgotten nutrient. Pp. 315–324 in Swine Nutrition, E.R. Miller, D.E. Ullrey and A.J. Lewis, eds.. Stoneham, MA: Butterworth-Heinemann. Vajrabukka, C., C. J. Thwaites, and D. J. Farrell. 1981. Overcoming the effects of high temperature on pig growth. Pp. 99–114 in Recent Advances in Animal Nutrition in Australia, D. J. Farrell and P. Vohra, eds. University of New England Publishing Unit, Armidale, Australia. Wahlstrom, R. C., A. R. Taylor, and R. W. Seerley. 1970. Effects of lysine in the drinking water of growing swine. J. Anim. Sci. 30: 368–373. Winks, W. R., A. K. Sutherland, and R. M. Salisbury. 1950. Nitrite poisoning of pigs. Queensl. J. Agric. Sci. 7:1–14. Wood, R. D., C. H. Chaney, D. G. Waddill, and G. W. Garrison. 1967. Effect of adding nitrate or nitrite to drinking water on the utilization of carotene by growing swine. J. Anim. Sci. 26: 510–513. Yang, T. S., B. Howard, and W. V. McFarlane. 1981. Effects of food on drinking behaviour of growing pigs. Appl. Anim. Ethol. 7: 259–270. Yang, T. S., M. A. Price, and F. X. Aherne. 1984. The effect of level of feeding on water turnover in growing pigs. Appl. Anim. Behav. Sci. 12:103–109.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

7 Nonnutrient feed additives are commonly included in swine diets. Of these, the antimicrobial agents are the additives most commonly used. Antimicrobial agents, along with anthelmintics, are defined as ‘‘drugs’’ by the Food and Drug Administration (FDA). Thus, their usage levels, allowable combinations, and periods of withdrawal prior to slaughter are regulated by the FDA and are published annually in the Feed Additive Compendium (1998). In addition, certain other additives are sometimes included in swine diets. The Association of American Feed Control Officials (1998) has established guidelines for the use of many of these products in animal feeds.

Nonnutritive Feed Additives

by Hays (1978), Zimmerman (1986), and Cromwell (1991). A summary of 1,194 experiments involving 32,555 pigs indicated that antimicrobials improved growth rate by 16.4 percent in weanling pigs (7 to 25 kg body weight), by 10.6 percent in growing pigs (17 to 49 kg), and by 4.2 percent in growing-finishing pigs (24 to 89 kg) (Hays, 1978; Zimmerman, 1986). Improvements in efficiency of feed utilization for these same groups were 6.9, 4.5, and 2.2 percent, respectively. Responses in pig growth to the feeding of antimicrobials are greater under field conditions than in controlled experiments at research stations (Cromwell, 1991). A summary of 67 field trials with young pigs indicated that the feeding of antimicrobials reduced mortality by one half (4.3 versus 2.0 percent), with even greater reductions in mortality when disease levels were high (15.6 versus 3.1 percent) (Maddox, 1985). Antibacterial agents also are effective in improving reproductive performance (Cromwell, 1991). A summary of nine experiments (1,931 sows) indicated that farrowing rate was improved from 75.4 percent in controls to 82.1 percent in treated sows, and the number of live pigs born was increased from 10.0 to 10.4, respectively, when antimicrobials were included in the diet at the time of breeding. In 11 experiments (2,105 sows), inclusion of antimicrobials in the lactation diet increased survival of pigs to weaning (84.9 versus 87.1 percent of pigs born alive) and pig weaning weights (4.65 versus 4.70 kg). Although the mechanism of action of antimicrobials is not well understood, their effects are generally grouped into three categories: a metabolic effect, a nutritional effect, and a disease-control effect. The first effect implies that these compounds directly influence certain metabolic processes in the animal (e.g., increased rate of protein synthesis). The second effect implies that antimicrobials cause changes in the microbial population that result in increased utilization of nutrients by the host animal. This effect is supported by evidence that antimicrobials reduce

ADDITIVES Antimicrobial Agents These are compounds that suppress or inhibit the growth of microorganisms. This class of compounds includes the antibiotics (naturally occurring substances produced by yeasts, molds, and other microorganisms) and the chemotherapeutics (chemically synthesized substances). They are added to feed at low (subtherapeutic) levels for growth promotion, improvement of feed utilization, reduction of mortality and morbidity, and improvement of reproductive performance. Antimicrobial agents also are used at moderate-to-high (prophylaxis) levels for the prevention of disease in exposed animals, and at high (therapeutic) levels for the treatment of certain swine diseases. Currently, 17 antimicrobial agents are approved for use in swine feed (Feed Additive Compendium, 1998). Of these, eight require withdrawal from the feed (on schedules ranging from 5 to 70 days) before animals are slaughtered, and nine do not require a withdrawal period. The efficacy of antimicrobials in improving the rate and efficiency of growth in pigs is well documented, as reviewed

97

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

98

Nutrient Requirements of Swine

intestinal wall thickness (thus improving absorption of nutrients), and that they reduce total gut mass (thus reducing heat loss from tissues with high metabolic activity). Most of the data support the disease control effect as the primary mode of action. This effect implies that antimicrobials suppress microorganisms that cause nonspecific, subclinical disease, thereby allowing the host animal to achieve a growth rate closer to its maximum potential. This suggested mechanism of action is supported by the greater response to antimicrobials that occurs in young versus older pigs, in a ‘‘dirty’’ versus ‘‘clean’’ environment, and in lowhealth versus high-health animals. Anthelmintics This class of drugs, also called ‘‘dewormers,’’ is included in feed to control internal parasites (e.g., roundworms, lungworms, threadworms). One drug, ivermectin, also is effective as a systemic for the control of external parasites (lice and mange). Of the seven anthelmintics currently approved for swine, five have specified withdrawal periods before slaughter (24 hours to 30 days) and two have no withdrawal period (Feed Additive Compendium, 1998). One compound in this group, dichlorvos, has been shown to reduce the incidence of stillbirths and increase pig weaning weights (Siers et al., 1976; Young et al., 1979) and may play a role in the immune response (Murray, 1983). Microbial Supplements Microbials that are directly fed, once referred to as ‘‘probiotics,’’ consist of live (viable), naturally occurring microorganisms such as Lactobacillus acidophilus, Streptococcus faecium, and Saccharomyces cerevisiae. The suggested action of these supplements is that they enhance the intestinal microbial balance in the host animal. In some instances, these products have been reported to benefit pig performance under field conditions, generally under high-stress conditions (Pollmann, 1986; Stavric and Kornegay, 1995); however, most controlled experiments at research stations have failed to show consistent, beneficial responses from their inclusion. A review of these products was written by van Belle et al. (1990). Oligosaccharides Inclusion of certain oligosaccharides (e.g., mannooligosaccharides, fructooligosaccharides) in the diet has been proposed to alter the ability of specific pathogens to colonize the intestinal tract (Monsand and Paul, 1995; Newman, 1995). The effect of oligosaccharides on performance of pigs is not well established. Some reports have shown a benefit in performance of young pigs from fructooligosaccharide inclusion (Hidaka et al., 1986; Fukuyasu and

Oshida, 1988), whereas others have not (Farnworth et al., 1991, 1992, 1995; Kornegay et al., 1992).

Enzymes Mixtures of cellulases, hemicellulases, and proteases are sometimes added to feeds in an attempt to improve the digestibility of complex carbohydrates and proteins. They are more commonly used in Europe, where diets are composed of a more diverse group of feedstuffs, than in North America, where diets tend to be based on corn or grain sorghum and soybean meal. Some research has shown these enzymes to be beneficial (Wenk, 1992). In areas where barley or rye is used, b-glucanase and pentosanases sometimes are included to degrade the b-glucans and pentosans (complex carbohydrates that interfere with digestibility of other nutrients) found in these cereal grains (Newman et al., 1980; Li et al., 1996), but improvements in pig performance do not necessarily occur (Thacker, 1993; Thacker and Baas, 1996). Varied responses have been shown to the addition of amylases and proteases to diets for very young pigs to aid in nutrient digestibility (Lewis et al., 1955; Cunningham and Brisson, 1957a,b; Combs et al., 1960). Recent reviews contain additional information on feed enzymes (Wenk and Boessinger, 1993; van Hartingsveldt et al., 1995). An enzyme that has recently received considerable attention is phytase. This enzyme cleaves the ortho-phosphate groups from phytic acid (phytate), the predominant form of phosphorus in cereal grains and oilseed meals. Phytase supplementation markedly improves the utilization of phytate phosphorus by pigs (Simons et al., 1990; Jongbloed et al., 1992; Cromwell et al., 1995) and reduces the excretion of phosphorus into the environment. For additional information on phytase, see Chapters 4 and 8.

Acidifiers Citric acid, fumaric acid, or formic acid additions to starter diets have been shown to enhance performance in early-weaned pigs (Kirchgessner and Roth, 1982, 1987; Falkowski and Aherne, 1984; Giesting and Easter, 1985; Burnell et al., 1988; Ravindran and Kornegay, 1993). Inorganic acids, such as phosphoric acid and, in some instances, hydrochloric acid, also have been found to be beneficial to young pig performance (Schoenherr, 1994; Bergstrom et al., 1996; Mahan et al., 1996). The mechanism of action is not clear, but it may be related to a reduction in pH in the upper intestinal tract, thereby reducing the potential for proliferation of undesirable microorganisms in the stomach and small intestine. Organic acids also have been used to preserve high-moisture grains and as mold inhibitors in feeds (Crenshaw et al., 1986).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nonnutritive Feed Additives

99

Flavors

Mineral Supplements

Synthetic flavors are added to feed to improve palatability and/or to mask off-flavors or off-odors in feed. Most research indicates that pigs may select diets with added flavors or aromatic compounds when given a choice; but when pigs are not given a choice, benefits from most flavors or aromatics are negligible (Hines, 1973; Hines et al., 1975; Kornegay et al., 1979; Ogunbameru et al., 1979). A review of flavors was written by McLaughlin et al. (1983).

High levels of dietary copper (100 to 250 ppm copper, as copper sulfate) have been shown to stimulate growth rate, feed intake, and efficiency of feed utilization in pigs, especially during the post-weaning and the early growth phases (Braude, 1945, 1975; Cromwell, 1991). Also, high dietary copper for sows has been found to increase pig weaning weights (Cromwell et al., 1993). Recent studies have also shown that high levels of zinc (3,000 ppm zinc, from zinc oxide) stimulate feed intake and growth rate in young pigs (Hahn and Baker, 1993; LeMieux et al., 1994; Hill et al., 1996). For further information and documentation, the reader is referred to the sections on copper and zinc in Chapter 4.

Odor Control Agents Sarsaponin, an extract from the yucca plant (Yucca schidigera), inhibits urease activity and is claimed to reduce odor in swine manure. In some instances, sarsaponin has been found to increase performance in weanling and growing-finishing pigs (Foster, 1983; Cromwell et al., 1985) and to reduce ammonia emissions (Sutton et al., 1992). Other products consisting of dried, live, naturally occurring microorganisms are claimed to reduce manure odor when added to feed. In some instances, zeolites have been shown to reduce odors and nitrogen volatilization (Barrington and El Moueddeb, 1995). Some of the oligosaccharides have been shown to alter hind gut microorganisms and reduce odor in swine manure (Sutton et al., 1991; Farnworth et al., 1995).

Antioxidants These products are added to feeds to inhibit oxidation of fat or vitamins. Examples are ethoxyquin and butylated hydroxytoluene (BHT).

Pellet Binders Certain clays (e.g., bentonite) are added to feed prior to pelleting in order to promote cohesiveness and inhibit crumbling of pellets. Some of the clays and zeolites also protect against aflatoxicosis in pigs by binding aflatoxins and preventing their absorption (Schell et al., 1993); however, they are not approved by the FDA for this purpose.

Flow Agents These products are the same as or similar to pellet binders. Their purpose is to prevent caking and improve the flow characteristics of certain ingredients. An example is hydrated sodium calcium aluminosilicate. Although not approved for aflatoxicosis prevention, this product is also effective in binding aflatoxins (Lindemann et al., 1993).

Carcass Modifiers Several b-adrenergic agonists, including clenbuterol, cimaterol, and ractopamine, increase carcass leanness when included in the diet (Jones et al., 1985; Moser et al., 1986; Cromwell et al., 1988; Watkins et al., 1990; Bark et al., 1992). However, these substances are not yet approved in the United States for use in swine. Under certain conditions, betaine and carnitine have been found to improve carcass leanness (Odle, 1995). Chromium also has been shown to improve carcass leanness when added to the diet as chromium picolinate in some instances (Page et al., 1993; Lindemann et al., 1995; Mooney and Cromwell, 1995), but not in others (Mooney and Cromwell, 1996; Crow and Newcomb, 1997). There is recent evidence that positional and geometric isomers of conjugated dienoic fatty acids (derivatives of linoleic acid [CLA]) reduce body fat and increase lean tissue when fed to mice, rats, and chicks (Pariza et al., 1996; Park et al., 1997), and though data are limited, CLA may produce a similar effect when fed to pigs (Pariza, 1997; Parrish et al., 1997). Certain carcass modifiers (e.g., b-agonists) can alter nutrient requirements (Anderson et al., 1987).

SAFETY CONCERNS There is concern by some that the feeding of antimicrobials to animals contributes to a reservoir of drug-resistant enteric bacteria that are capable of transferring their resistance to pathogenic bacteria, thereby causing a potential public health hazard (Smith, 1962; Falkow, 1975; Linton, 1977). The greatest concern is in regard to penicillin and the tetracyclines, because they also are used in human medicine. Although transfer of antibiotic resistant plasmids (Rplasmids) occurs in vitro, the extent to which it occurs in the animal, and between animal bacteria and human

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

100

Nutrient Requirements of Swine

bacteria, is not well documented. Animal bacteria do not colonize very effectively in humans unless extremely large doses are consumed; and even then, they are transient (Smith, 1969). In 1987, the Food and Drug Administration asked the Institute of Medicine of the National Academy of Sciences to conduct an independent review of the human health consequences and make a quantitative risk assessment associated with the use of penicillin and the tetracyclines at subtherapeutic levels in animal feeds. The committee was unable to find a substantive body of direct evidence that established the existence of a definite health hazard in humans associated with the use of these antimicrobials in animal feeds (Institute of Medicine, 1988). Similarly, other task forces concluded that there was no conclusive evidence of human health being compromised by subtherapeutic antimicrobial usage in animals (National Research Council, 1980; Council for Agricultural Science and Technology, 1981). Monitoring and surveillance of bacterial resistance in animals and in humans has continued, with no animal-tohuman infection path being clearly delineated. Although the incidence of antimicrobial resistance in the human population remains high, there is no evidence that the levels or patterns have changed (Lorian, 1986). Although antimicrobial agents have been fed to billions of animals for over 45 years, there is still no convincing evidence of any unfavorable health effects in humans that can be directly linked to the feeding of subtherapeutic levels of antibiotics to animals.

REGULATIONS Regulations and constraints on the use of feed additives vary among countries. In addition, the approved uses of additives are subject to change. For official information concerning U.S. Food and Drug Administration approval of feed additives and other animal drugs, the Code of Federal Regulations, Title 21, should be consulted. A revised edition of Title 21 is available in April of each year. Individual issues of the Federal Register update the Code of Federal Regulations. The Federal Register and the Code of Federal Regulations must be used together to determine the latest version of any rule. Title 21 is published in six parts. Part 500-599 covers animal drugs, feeds, and related products. It is available from the Superintendent of Documents, U.S. Government Printing Office, Washington, DC 20402. The Federal Register is available from the same address and includes monthly issues of the ‘‘List of Code of Federal Regulations Sections Affected’’ and ‘‘The Federal Register Index.’’ Additional information on feed additives, usage levels, and legal requirements is available in the Feed Additive

Compendium, which is published yearly by the Miller Publishing Company, 12400 Whitewater Drive, Minneapolis, MN 55343, and in the compendium of Medicating Ingredient Brochures, Feed and Fertilizer Division, published by Agriculture Canada, Ottawa, Ontario, Canada. The Compendium of Medicating Ingredient Brochures is available from Supply and Services Canada, Canada Communications Group, Ottawa, Ontario, Canada, K1A 0S9.

REFERENCES Anderson, D. B., E. L. Veenhuizen, W. P. Waitt, R. E. Paxton, and S. S. Young. 1987. The effect of dietary protein on nitrogen metabolism, growth performance and carcass composition of finishing pigs fed ractopamine. Fed. Proc. 46:1021 (Abstr.). Association of American Feed Control Officials. 1998. AAFCO Official Publication. Atlanta: Georgia Dept. of Agric. Bark, L. J., T. S. Stahly, G. L. Cromwell, and J. Miyat. 1992. Influence of genetic capacity for lean tissue growth on rate and efficiency of tissue accretion in pigs fed ractopamine. J. Anim. Sci. 70:3391–3400. Barrington, S., and K. El Moueddeb. 1995. Zeolite to control swine manure odours and nitrogen volatilization. Pp. 65–68 in Proc. International Livestock Odor Conference. Ames: Iowa State University. Bergstrom, J. R., J. L. Nelssen, and M. D. Tokach, and R. D. Goodband. 1996. An evaluation of several diet acidifiers commonly used in pig starter diets to improve growth performance. J. Anim. Sci. 74:194 (Abstr.). Braude, R. 1945. Some observations on the need for copper in the diet of fattening pigs. J. Agric. Sci. 35:163–167. Braude, R. 1975. Copper as a performance promoter in pigs. Pp. 79–97 in Farming Symp. Development Assoc., London. Burnell, T. W., G. L. Cromwell, and T. S. Stahly. 1988. Effects of dried whey and copper sulfate on the growth responses to organic acid in diets for weanling pigs. J. Anim. Sci. 66:1100–1108. Combs, G. E., W. L. Alsmeyer, H. D. Wallace, and M. Koger. 1960. Enzyme supplementation of baby pig rations containing different sources of carbohydrate and protein. J. Anim. Sci. 19:932–937. Council for Agricultural Science and Technology. 1981. Antibiotics in Animal Feeds. Report No. 88. Ames, IA: Council for Agricultural Science and Technology. Crenshaw, J. D., E. R. Peo, Jr., A. J. Lewis, and N. R. Schneider. 1986. The effects of sorbic acid in high moisture sorghum grain diets on performance of weanling swine. J. Anim. Sci. 63:831–837. Cromwell, G. L. 1991. Antimicrobial agents. Pp. 297–314 in Swine Nutrition, E. R. Miller, D. E. Ullrey, and A. J. Lewis, eds. Stoneham, MA: Butterworth-Heinemann. Cromwell, G. L., T. S. Stahly, and H. J. Monegue. 1985. Efficacy of sarsaponin for weanling and growing-finishing swine housed at two animal densities. J. Anim. Sci. 61(Suppl. 1):111 (Abstr.). Cromwell, G. L., J. D. Kemp, T. S. Stahly, and R. H. Dalrymple. 1988. Effects of dietary level and withdrawal time on the efficacy of cimaterol as a growth repartitioning agent in finishing swine. J. Anim. Sci. 66:2193–2199. Cromwell, G. L., H. J. Monegue, and T. S. Stahly. 1993. Long-term effects of feeding a high copper diet to sows during gestation and lactation. J. Anim. Sci. 71:2996–3002. Cromwell, G. L., R. D. Coffey, G. R. Parker, H. J. Monegue, and J. H. Randolph. 1995. Efficacy of a recombinant-derived phytase in improving the bioavailability of phosphorus in corn–soybean meal diets for pigs. J. Anim. Sci. 73:2000–2008.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nonnutritive Feed Additives Crow, S. D., and M. D. Newcomb. 1997. Effect of dietary chromium additions along with varying protein levels on growth performance and carcass characteristics. J. Anim. Sci. 74:79 (Abstr.). Cunningham, H. M., and G. J. Brisson. 1957a. The effect of amylases on the digestibility of starch by baby pigs. J. Anim. Sci. 16:370–376. Cunningham, H. M., and G. J. Brisson. 1957b. The effect of proteolytic enzymes on the utilization of animal and plant proteins by newborn pigs and the response to predigested protein. J. Anim. Sci. 16:568–573. Falkow, S. 1975. Infectious Multiple Drug Resistance. London: Pion Ltd. Falkowski, J. F., and F. X. Aherne. 1984. Fumaric and citric acid as feed additives in starter pig nutrition. J. Anim. Sci. 58:935–938. Farnworth, E. R., N. Dilawri, H. Yamazaki, H. W. Modler, and J. D. Jones. 1991. Studies on the effect of adding Jerusalem artichoke flour to pig milk replacer. Can. J. Anim. Sci. 71:531–536. Farnworth, E. R., H. W. Modler, J. D. Jones, N. Cave, H. Yamazaki, and A. V. Rao. 1992. Feeding Jerusalem artichoke flour rich in fructooligosaccharides to weanling pigs. Can. J. Anim. Sci. 72:977–980. Farnworth, E. R., H. W. Modler, and D. A. Mackie. 1995. Adding Jerusalem artichoke (Helianthus tuberosus L.) to weanling pig diets and the effect on manure composition and characteristics. Anim. Feed Sci. Tech. 55:153–160. Feed Additive Compendium. 1998. Minneapolis, Minn.: Miller Publishing Co. Foster, J. R. 1983. Effects of sarsaponin in growing-finishing swine diets. J. Anim. Sci. 57(Suppl. 1):94 (Abstr.). Fukuyasu, T., and Oshida, T. 1988. Use of Neosugart in piglets. P. 1 in Proc. 3rd Neosugart Conf., Tokyo, Japan, 1986. Giesting, D. W., and R. A. Easter. 1985. Response of starter pigs to supplementation of corn–soybean meal diets with organic acids. J. Anim. Sci. 60:1288–1294. Hahn, J. D., and D. H. Baker. 1993. Growth and plasma zinc responses of young pigs fed pharmacologic levels of zinc. J. Anim. Sci. 71:3020–3024. Hays, V. W. 1978. Effectiveness of Feed Additive Usage of Antibacterial Agents in Swine and Poultry Production. Report to the Office of Technology Assessment, U.S. Congress. U.S. Government Printing Office, Washington, D.C. (Edited version: Hays, V.W., 1981. The Hays Report. Rachelle Laboratories, Inc., Long Beach, CA.) Hidaka, H., T. Eida, T. Takizawa, T. Tokunaga, and Y. Tashiro. 1986. Effects of fructooligosaccharides on intestinal flora and human health. Bifidobacteria Microflora 5:37–50. Hill, G. M., G. L. Cromwell, T. D. Crenshaw, R. C. Ewan, K. A. Knabe, A. J. Lewis, D. C. Mahan, G. C. Shurson, L. L. Southern, and T. L. Veum, NCR-42 and S-145 Regional Swine Nutrition Committees. 1996. Impact of pharmacological intakes of zinc and(or) copper on performance of weanling pigs. J. Anim. Sci. 74(Suppl. 1):181 (Abstr.). Hines, R. H. 1973. Feed flavors in swine starter rations. Pp. 37–42 in Proc. Swine Industry Day, Kansas State Univ. Hines, R. H., B. A. Koch, and G. L. Allee. 1975. Attractants for swine starter feed: Aroma vs. taste. Pp. 20–23 in Proc. Swine Industry Day, Kansas State Univ. Institute of Medicine. 1988. Human Health Risks with the Subtherapeutic Use of Penicillin or Tetracycline in Animal Feed. Institute of Medicine, National Academy of Sciences. Washington, DC: National Academy Press. Jones, R. W., R. A. Easter, F. K. McKeith, R. H. Dalrymple, H. M. Maddock, and P. J. Bechtel. 1985. Effect of the b-adrenergic agonist cimaterol (CL 263,780) on the growth and carcass characteristics of finishing swine. J. Anim. Sci. 61:905–913. Jongbloed, A.W., Z. Mroz, and P. A. Kemme. 1992. The effect of supplementary Aspergillus niger phytase in diets for pigs on concentration and apparent digestibility of dry matter, total phosphorus, and phytic acid in different sections of the alimentary tract. J. Anim. Sci. 70:1159–1168.

101

Kirchgessner, M., and F. X. Roth. 1982. Fumaric acid as a feed additive in pig nutrition. Pig News and Information 3:259–263. Kirchgessner, M., and F. X. Roth. 1987. Use of formates in the feeding of piglets. First communication: Calcium formate. Landwirtsch. Forsch. 40:141–152. Kornegay, E. T., S. E. Tinsley, and K. L. Bryant. 1979. Evaluation of rearing systems and feed flavors for pigs weaned at two to three weeks of age. J. Anim. Sci. 48:999–1006. Kornegay, E. T., C. M. Wood, and L. A. Eng. 1992. Effectiveness and safety of fructooligosaccharides for pigs. J. Anim. Sci. 70(Suppl. 1):19 (Abstr.). LeMieux, F. M., L. L. Southern, and T. D. Bidner. 1994. Effect of dietary zinc and(or) antibiotic on growth of pigs weaned at four weeks of age. J. Anim. Sci. 72(Suppl. 2):6 (Abstr.). Lewis, C. J., D. V. Catron, C. H. Liu, V. C. Speer, and G. C. Ashton. 1955. Enzyme supplementation of baby pig diets. J. Agric. Food Chem. 3:1047–1050. Li, S., W. C. Sauer, R. Mosenthin, and B. Kerr. 1995. Effect of bglucanase supplementation of cereal-based diets for starter pigs on the apparent digestibilities of dry matter, crude protein, and energy. Anim. Feed Sci. Tech. 59:223–231. Lindemann, M. D., D. J. Blodgett, E. T. Kornegay, and G. G. Schurig. 1993. Potential ameliorators of aflatoxicosis in weanling/growing swine. J. Anim. Sci. 71:171–178. Lindemann, M. D., C. M. Wood, A. F. Harper, E. T. Kornegay, and R. A. Anderson. 1995. Dietary chromium picolinate additions improve gain:feed and carcass characteristics in growing-finishing pigs and increase litter size in reproducing sows. J. Anim. Sci. 73:457–465. Linton, A. H. 1977. Antibiotics, animals and man—an appraisal of a contentious subject. Pp. 315–343 in Antibiotics and Antibiosis in Agriculture, M. Woodbine, ed. Woburn, MA: Butterworths. Lorian, V. 1986. Antibiotic sensitivity patterns of human pathogens in American hospitals. J. Anim. Sci. 62(Suppl. 3):49–55. Maddox, H. M. 1985. Unpublished data from American Cyanamid Co., Princeton, NJ (cited by Cromwell, 1991). Mahan, D. C., E. A. Newton, and K. R. Cera. 1996. Effect of supplemental sodium chloride, sodium phosphate, or hydrochloric acid in starter pig diets containing dried whey. J. Anim. Sci. 74:1218–1222. McLaughlin, C. L., C. A. Baile, L. L. Buckholtz, and S. K. Freeman. 1983. Preferred flavors and performance of weanling pigs. J. Anim. Sci. 56:1287–1293. Monsand, P. F., and F. Paul. 1995. Oligosaccharide feed additives. Pp. 233–245 in Biotechnology in Animal Feeds and Animal Feeding, R. J. Wallace and A. Chesson, eds. Weinheim, Germany: VCH Verlagsgesellschaft mbH. Mooney, K. W., and G. L. Cromwell. 1995. Effects of dietary chromium picolinate supplementation on growth, carcass characteristics, and accretion rates of carcass tissues in growing-finishing swine. J. Anim. Sci. 75:3351–3357. Mooney, K. W., and G. L. Cromwell. 1996. Effects of chromium picolinate on performance and tissue accretion in pigs with different lean gain potential. J. Anim. Sci. 74 (Suppl. 1): 65 (Abstr.). Moser, R. L., R. H. Dalrymple, S. G. Cornelius, J. E. Pettigrew, and C. E. Allen. 1986. Effect of cimaterol (CL 263,780) as a repartitioning agent in the diet for finishing pigs. J. Anim. Sci. 62:21–26. Murray, F. A. 1983. Effects of dichlorvos on lymphocyte reactivity during pregnancy in the pig. J. Anim. Sci. 57:1270–1275. National Research Council. 1980. Effects on Human Health of Subtherapeutic Use of Antimicrobials in Animal Feeds. Washington, DC: National Academy Press. 376 pp. Newman, C. W., R. F. Eslick, J. W. Pepper, and A. M. El-Negoumy. 1980. Performance of pigs fed hulless and covered barleys supplemented with or without a bacterial diastase. Nutr. Rep. Internat. 22:833–837.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

102

Nutrient Requirements of Swine

Newman, K. E. 1995. Mannan oligosaccharides: Immune modulator or rumen efficiency potentiator. Pp. 37–42 in Proc. 56th Minnesota Nutrition Conference Alltech Tech. Symposium. St. Paul: University of Minnesota. Odle, J. 1995. Betaine and carnitine—evaluation of performance and carcass effects. Pp. 1–14 in Proc. Carolina Swine Nutr. Conference. Raleigh: North Carolina State Univ. Ogunbameru, B. O., E. T. Kornegay, K. L. Bryant, K. H. Hinkelmann, and J. W. Knight. 1979. Evaluation of a fed flavour in lactation and starter diets to stimulate feed intake of weaned pigs. Nutr. Rep. Intl. 20:455–460. Page, T. G., L. L. Southern, T. L. Ward, and D. L. Thompson, Jr. 1993. Effect of chromium picolinate on growth and serum- and carcass traits of growing-finishing pigs. J. Anim. Sci. 71:656–662. Pariza, M. W. 1997. Conjugated linoleic acid, a newly recognized nutrient. Chem. Ind. 12:464–466. Pariza, M., Y. Park, M. Cook, K. Albright, and W. Liu. 1996. Conjugated linoleic acid (CLA) reduces body fat. FASEB Journal 10:A560 (Abstr.). Park, Y., K. J. Albright, W. Liu, J. M. Storkson, M. E. Cook, and M. W. Pariza. 1997. Effect of conjugated linoleic acid on body composition in mice. Lipids. 32:853–858. Parrish, F.C., Jr., R.L. Thiel, J. C. Sparks, and R. C. Ewan. 1998. Effects of conjugated linoleic acid (CLA) on swine performance and body composition. 1997 Swine Research Report, Iowa State University, AS-638:187–190. Pollmann, D. S. 1986. Probiotics in pig diets. Pp. 193–205 in Recent Advances in Animal Nutrition, W. Haresign and D. J. A. Cole, eds. London: Butterworths. Ravindran, V., and E. T. Kornegay. 1993. Acidification of weaner pig diets: A review. J. Anim. Sci. Food Agric. 62:313–322. Schell, T. C., M. D. Lindemann, E. T. Kornegay, D. J. Blodgett, and J. A. Doerr. 1993. Effectiveness of different types of clay for reducing the detrimental effects of aflatoxin-contaminated diets on performance and serum profiles of weanling pigs. J. Anim. Sci 71:1226–1231. Schoenherr, W. D. 1994. Phosphoric acid-based acidifiers explored for starter diets. Feedstuffs, September 26, 1994, p. 14. Siers, D. G., D. E. DeKay, H. J. Mersmann, L. J. Brown, and H. C. Stanton. 1976. Late gestation feeding of dichlorvos: a physiological characterization of the neonate and a growth-survival response. J. Anim. Sci. 42:381–392. Simons, P. C. M., H. A. J. Versteegh, A. W. Jongbloed, P. A. Kemme, P. Slump, K. D. Bos, M. G. E. Wolters, R. F. Beudeker, and G. J. Verschoor. 1990. Improvement of phosphorus availability by microbial phytase in broilers and pigs. Br. J. Nutr. 64:525–540.

Smith, H. W. 1962. The effects of the use of antibiotics on the emergence of antibiotic-resistant disease-producing organisms in animals. Pp. 374–388 in Antibiotics in Agriculture. Univ. of Nottingham Ninth Easter School in Agriculture Science. London: Butterworths. Smith, H. W. 1969. Transfer of antibiotic resistance from animal and human strains of Escherichia coli to resistant E. coli in the alimentary tract of man. Lancet 1:1174–1176. Stavric, S., and E. T. Kornegay. 1995. Microbial probiotics for pigs and poultry. Pp. 205–231 in Biotechnology in Animal Feeds and Feeding, R. J. Wallace and A. Chesson, eds. Weinheim, Germany: VCH Verlagsgesellschaft. Sutton, A. L., A. G. Mathew, A. B. Scheidt, J. A. Patterson, and D. T. Kelly. 1991. Effect of carbohydrate source and organic acids on intestinal microflora and performance of the weanling pig. Pp. 422–427 in Proc. 5th Congress on Digestive Physiology in Pigs. EEAP Pub. No. 54. Pudoc, Wageningen, Netherlands. Sutton, A. L., S. R. Goodall, J. A. Patterson, A. G. Mathew, D. T. Kelly, and K. A. Meyerholtz. 1992. Effects of odor control compounds on urease activity in swine manure. J. Anim. Sci. 70(Suppl. 1):160 (Abstr.). Thacker, P. A. 1993. Novel approaches to growth promotion in the pig. Pp. 295–306 in Recent Developments in Pig Nutrition, D. J. A. Cole, W. Haresign, and P. C. Garnsworthy, eds. Nottingham, U.K.: Nottingham University Press. Thacker, P.A., and F. C. Baas. 1996. Effects of gastric pH on the activity of exogenous pentosanase and the effect of pentosanase supplementation of the diet on the performance of growing-finishing pigs. Anim. Feed Sci. Tech. 63:187–200. van Belle, M., E. Teller, and M. Focant. 1990. Probiotics in animal nutrition: A review. Arch. Anim. Nutr., Berlin 40:7:543–567. van Hartingsveldt, W., M. Hessing, J. P. van der Lugt, and W.A.C. Somers. 1995. The Second European Symposium on Feed Enzymes. Zeist, Netherlands: TNO Nutrition and Food Research Institute. 302 pp. Watkins, L. E., D. J. Jones, D. H. Mowrey, D. B. Anderson, and E. L. Veenhuizen. 1990. The effects of various levels of ractopamine hydrochloride on the performance and carcass characteristics of finishing swine. J. Anim. Sci. 68:3588–3595. Wenk, C. 1992. Enzymes in the nutrition of monogastric farm animals. Pp. 205–218 in Biotechnology in the Feed Industry, T. P. Lyons, ed. Nicholasville, KY: Alltech Technical Publications. Wenk, C., and M. Boessinger. 1993. Enzymes in Animal Nutrition. Zurich, ¨ Switzerland: Institut fur Nutztierwissenschaften, Gruppe Ernahrung. Young, R., Jr., D. K. Hass, and L. J. Brown. 1979. Effect of late gestation feeding of dichlorvos in non-parasitized and parasitized sows. J. Anim. Sci. 48:45–51. Zimmerman, D. R. 1986. Role of subtherapeutic antimicrobials in pig production. J. Anim. Sci. 62(Suppl. 3):6.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

8 Maximization of individual pig performance traditionally has been the goal of swine producers and nutritionists. Diets are generally formulated to achieve this goal with little or no regard for the amount of nutrients excreted. Consequently, oversupplementation of diets with nutrients to ensure maximum pig performance results in excessive amounts of excreted nutrients in feces and urine. Because of its relatively high content of nitrogen, phosphorus, potassium, and other nutrients, manure is an excellent fertilizer when applied to land. During the past decade, there has been little change in total number of hogs produced in the United States, but the number of large swine units and the intensity of production has sharply increased. As a result, large amounts of manure are produced on a much smaller land area. Distribution and disposal has become a problem. The application of excessive amounts of manure to land can potentially lead to surface and ground water contamination and to the accumulation of minerals in the soil. Of the nutrients present in manure, nitrogen, phosphorus, sodium, potassium, copper, and zinc cause the greatest concern. In some areas, nitrogen is used as the basis to regulate the amount of manure that can be applied to the land. However, evidence is accumulating which suggests that phosphorus will be the nutrient that limits land application of manure in the more intensive swine producing areas. For example, Barker and Zublena (1995) reported that in North Carolina, animal manure could provide about 20 percent of the nitrogen and 66 percent of the phosphorus requirements of all non-legume agronomic crops and forages produced in that state. Three of the 100 counties surveyed had enough manure to exceed the nitrogen requirements of their crops, and 18 counties had enough manure to exceed the phosphorus requirements of crops. Soil analyses of a Sampson County (North Carolina) bermudagrass pasture that was fertilized with swine lagoon effluent to satisfy nitrogen requirements showed approxi-

Minimizing Nutrient Excretion

mately a fourfold increase in phosphorus and zinc, a onefold increase in potassium, and a threefold increase in copper to a depth of 91 cm during the three-year period of application (Mueller et al., 1994). These findings may well be representative of other regions of the United States. The overall quality of water, both surface water and ground water, can be negatively affected by applying excess nitrogen and phosphorus, and perhaps other nutrients, to soil. Excess nitrogen application can lead to increases in nitrate content of ground water and to potential runoff of nitrate into surface water. Excess phosphorus application results in excess buildup of phosphorus in the soil. While phosphorus is adsorbed onto soil particles and does not leach into ground water, it can erode (along with soil particles) into streams, lakes, and rivers. Phosphorus is the most limiting nutrient that regulates aquatic plant growth (Pierzynski et al., 1994; Sharpley et al., 1994), so when it is added to bodies of surface water, phosphorus stimulates growth of algae and other aquatic vegetation. Decomposition of such vegetation can lead to a general deterioration of water quality, a process called ‘‘eutrophication’’ (Crenshaw and Johanson, 1995). To prevent potential pollution by nitrogen and phosphorus, governments in many countries have passed legislation requiring nutrient management plans for each farm, so that land application of manure can be properly controlled (Hacker and Du, 1993). Even in the best situation, pigs do not utilize 100 percent of nutrients consumed. A review of balance data for pigs fed commercial feedstuffs indicates the following apparent utilization values (as percent of intake): 30 to 55 for nitrogen; 30 to 50 for calcium; 20 to 50 for phosphorus; 5 to 20 for potassium; 10 to 25 for sodium; 15 to 30 for magnesium; 5 to 30 for copper; 5 to 30 for zinc; 5 to 10 for manganese; and 5 to 30 for iron (Kornegay and Harper, 1997). Therefore, the percentages of intake excreted are 45 to 60 percent of nitrogen; 50 to 80 percent of calcium and phosphorus; and 70 to 95 percent of potassium,

103

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

104

Nutrient Requirements of Swine

sodium, magnesium, copper, zinc, manganese, and iron. The amount of a nutrient excreted can be influenced by several factors including quality, source, and level of the nutrient fed; the level and proportion of other nutrients; processing methods; age, class, and nutritional status of animals; and environmental factors. The use of highly digestible feedstuffs in diets is an effective means of reducing excretion of nitrogen and other nutrients. A portion of the nutrients excreted by the pig is a direct result of feeding excessive levels of nutrients. Results of surveys of the nutrient composition of diets indicate that diets commonly include excessive amounts of certain nutrients. Nutritionists call these excesses a safety factor included in the diet to allow for the variability of nutrient composition of feed ingredients or to compensate for uncertainty about the availability of the nutrients. Results of a survey by Cromwell (1989a,b) of phosphorus recommendations of several universities and feed companies showed that the average range of university recommendations was 110 to 120 percent of National Research Council (1988) requirements, whereas the average range of industry recommendations was 120 to 130 percent of these requirements. Spears (1996) reported that the mineral concentrations of sow and finishing pig diets analyzed by the North Carolina Feed Testing Laboratory were greatly in excess of requirements. The median levels as a percentage of National Research Council (1988) requirements for sow and finishing pig diets were the following: 161 and 192 percent, respectively, for calcium; 140 and 155 percent for phosphorus; 147 and 190 percent for sodium; 390 and 423 percent for potassium; 525 and 400 percent for magnesium; 440 and 667 percent for copper; 470 and 776 percent for iron; 770 and 3,100 percent for manganese; and 334 and 298 percent for zinc. Other surveys have reported similar findings of diets containing excess levels of nutrients. Excretion of minerals could be markedly reduced simply by reducing these excessive levels of nutrients in diets. Using high-quality protein sources with superior amino acid balance and formulating diets to achieve an ideal protein basis reduces nitrogen excretion. Lowering the dietary protein level and supplementing with certain crystalline amino acids also reduce nitrogen excretion. The reason is that both procedures reduce excesses of unneeded amino acids, which otherwise are degraded and excreted as urea nitrogen. Bridges et al. (1994) and Carter et al. (1996) showed that nitrogen excretion could be reduced by 30 to 40 percent by feeding corn–soybean meal diets in which the protein level was reduced by 4 percentage points and the diets supplemented with lysine, threonine, tryptophan, and methionine. Kerr and Easter (1995) suggested that for each one percentage unit reduction in dietary crude protein combined with amino acid supplementation, total nitrogen losses (fecal and urinary) could be reduced by approximately 8 percent. Conversely,

the use of low-quality protein sources (e.g., hydrolyzed hog hair meal) markedly increases nitrogen excretion (Kornegay, 1978b). Also, the inclusion of high levels of crude fiber in the diet reduces the efficiency of nitrogen utilization (Kornegay, 1978a). In corn–soybean meal diets, two-thirds of the phosphorus is bound as phytic acid and is poorly available to the pig (Cromwell and Coffey, 1991); hence, much of the phosphorus is excreted. The amount excreted can be significantly decreased by the inclusion of microbial phytase in the diet, which releases some of the bound phosphorus, making it available to the pig (Jongbloed et al., 1992; Cromwell et al., 1993). Thus, the amount of inorganic phosphorus that must be added to meet the available phosphorus requirement is reduced, and phosphorus excretion can be decreased by 30 to 50 percent (Bridges et al., 1995; Carter et al., 1996). The magnitude of the response to microbial phytase has been shown to be influenced by the source of phosphorus, dietary level of available phosphorus, the amount of phytase added, and the ratio of calcium to phosphorus (Lei et al., 1994; Kornegay, 1996). Microbial phytase also releases calcium (Mroz et al., 1994; Radcliffe et al., 1995), zinc (Lei et al., 1993; Pallauf et al., 1994), as well as some amino acids (Kemme et al., 1995) that may be bound by phytic acid. High dietary levels of copper and zinc also significantly increase the amount of copper and zinc that is excreted. In a study by Apgar and Kornegay (1996), 71-kg barrows excreted 6.7 times more copper when fed diets containing 218 versus 32 ppm copper. Calculations based on data reported by Adeola et al. (1995) for 15- to 18-kg pigs fed diets with 23 or 123 ppm zinc from ZnSO4 indicated that pigs fed the low-zinc diet excreted 16 mg of zinc per day, whereas pigs fed the high-zinc diet excreted 61 mg of zinc per day, a 3.8-fold increase in the amount of zinc excreted. When diets containing 2,500 to 3,000 ppm zinc are fed to weanling pigs as is commonly done for growth promotion (Hahn and Baker, 1993; LeMieux et al., 1995; Smith et al., 1995; Hill et al., 1996), approximately 90 to 95 percent of the zinc will be excreted. Although these high levels would be fed for a period of only a few weeks, the total amount of zinc excreted could approach or exceed the total amount of zinc excreted during the entire growing-finishing period by pigs fed diets containing approximately 100 ppm zinc. Other strategies also have potential for reducing nutrients excreted. For example, improvements in overall feed efficiency can produce a reduction in excreted nutrients. Henry and Dourmad (1992) reported for growing-finishing pigs that for each 0.1 percentage unit decrease in feedto-gain ratio, there was a 3 percent decrease in nitrogen excreted. Improvements in feed efficiency could result from improved genetics, improved environmental conditions, proper formulation of diets using high-quality ingre-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Minimizing Nutrient Excretion dients, use of pelleting and fine grinding of feed, and proper feeder adjustment to reduce wastage. Harper (1994) estimated that a 5 percent level of feed waste resulted in an additional 327 g of nitrogen and 82 g of phosphorus excreted per animal. Nutrient requirements change as pigs increase in body weight. Thus, frequent changes in diet formulation can meet the nutrient needs of the pig more efficiently. Frequent adjustments in diets can result in reduced intake of nutrients and, thus, reduced excretion of nutrients. Phase feeding and separate-sex feeding are ways to meet the nutrient needs of growing and finishing pigs more precisely and reduce nitrogen excretion. A further point is that the efficiency of animal performance follows the principle of diminishing returns in response to nutrient input (Heady et al., 1954; Combs et al., 1991; Gahl et al., 1995). Heady et al. (1954) reported that in 14 of 16 years, swine diets formulated using the diminishing return concept would have produced greater profits than diets formulated for maximum gain. As the cost of disposing of nitrogen and phosphorus increases, the nutrient levels fed to pigs will probably decrease. In the future, nutritionists may formulate diets to achieve 95 to 98 percent rather than 100 percent of maximum response, because the benefit of adding a unit of nutrient increases at a decreasing rate, and nutrient costs increase at an increasing rate as the animal reaches maximum performance. The success of all strategies for reducing nutrients excreted is dependent on an accurate estimate of the nutrient requirements of the class of pigs in question and on the accuracy of compositional information for, and bioavailability of, feed ingredients. Recommended nutrient requirements for the different classes of pigs often vary and, in many cases, are only estimates for an ‘‘average’’ animal under ‘‘average’’ environmental conditions. The estimated nutrient requirements may be influenced by the animal’s genetic potential, feeding methods, environmental conditions, the ingredients used, and animal response criteria. With the exception of phosphorus and amino acids, nutrient requirements are generally based on total nutrients rather than available nutrients. The available nutrient requirement of pigs can be accurately met, assuming they are known, only if the compositional data of feed ingredients are expressed on available nutrient composition. Using more precise data on compositional and nutrient availability for feed ingredients and better defined requirements will allow nutritionists to formulate diets that more precisely meet the needs of the animal at the various stages of production. In summary, nutrients excreted can be reduced through proper nutrient management to improve the availability of nutrients and reduce excessive amounts that are fed. In the future, diet formulation will be integrated into a total

105

production system with nutrient and manure management being its major components. The need for more careful nutrient management planning probably will increase in the future, as the intensity of the industry increases and as the concerns of the public increase.

REFERENCES Adeola, O., B. V. Lawrence, A. L. Sutton, and T. R. Cline. 1995. Phytaseinduced changes in mineral utilization in zinc-supplemented diets for pigs. J. Anim. Sci. 73:3384–3391. Apgar, G. A., and E. T. Kornegay. 1996. Mineral balance of finishing pigs fed copper sulfate or a copper lysine complex at growth stimulating levels. J. Anim. Sci. 74:1594–1600. Barker, J. C., and J. P. Zublena. 1995. Livestock manure nutrient assessment in North Carolina. Pp. 98–106 in Proc. Seventh International Symposium on Agricultural and Food Processing Wastes. Sponsored by ASAE, Chicago, IL, June 18–20. Bridges, T. C., L. W. Turner, G. L. Cromwell, and J. L. Pierce. 1995. Modeling the effects of diet formulation on nitrogen and phosphorus excretion in swine waste. Applied Engineering in Agriculture 11(5):731–739. Carter, S. D., G. L. Cromwell, M. D. Lindemann, L. W. Turner, and T. C. Bridges. 1996. Reducing N and P excretion by dietary manipulation in growing and finishing pigs. J. Anim. Sci. 74(Suppl. 1):59 (Abstr.). Combs, N. R., E. T. Kornegay, M. D. Lindemann, and D. R. Notter. 1991. Calcium and phosphorus requirement of swine from weaning to market weight: 1. Development of response curves for performance. J. Anim. Sci. 69:673–681. Crenshaw, T. D., and J. C. Johanson. 1995. Nutritional strategies for waste reduction management: Minerals. Pp. 69–78 in New Horizons In Animal Nutrition and Health; J. B. Longenecker and J. W. Spears, eds. The Institute of Nutrition of The University of North Carolina, Chapel Hill, Nov. 7 and 8. Cromwell, G. L. 1989a. Requirements, biological availability of calcium, phosphorus for swine evaluated. Feedstuffs 60(23):16. Cromwell, G. L. 1989b. Requirements and biological availability of phosphorus for swine. Pp. 75–95 in Proc. Pitman-Moore Nutr. Conf., Des Moines, IA. Cromwell, G. L., and R. D. Coffey. 1991. Phosphorus—a key essential nutrient, yet a possible major pollutant—its central role in animal nutrition. Pp. 133–145 in Biotechnology in the Feed Industry, T. P. Lyons, ed. Nicholasville, KY: Alltech Technical Publications. Cromwell, G. L., T. S. Stahly, R. D. Coffey, H. J. Monegue, and J. H. Randolph. 1993. Efficacy of phytase in improving the bioavailability of phosphorus in soybean meal and corn–soybean meal diets for pigs. J. Anim. Sci. 71:1831–1840. Gahl, M. J., T. D. Crenshaw, and N. J. Benevenga. 1995. Diminishing returns in weight, nitrogen, and lysine gain of pigs fed six levels of lysine from three supplemental sources. J. Anim. Sci. 72:3177–3187. Hacker, R. R., and Z. Du. 1993. Livestock pollution and politics. Pp. 3–21 in Nitrogen flow in pig production and environmental consequences; M. W. A. Verstegen, L. A. den Hartog, G. J. M. van Kempen and C. J. H. M. Metz, eds. EAAP publ. 69, Pudoc Scientific Publishers, Wageningen. Hahn, J. D., and D. H. Baker. 1993. Growth and plasma zinc responses of young pigs fed pharmacologic levels of zinc. J. Anim. Sci. 71:3020–3024. Harper, A. F. 1994. Feeding technologies to reduce excess nutrients in swine diets. Pp. 44–51 in Proc. Meeting the Challenge of Environmental Management on Hog Farms. Second Annual Virginia Tech Swine Producers Seminar, Carson, Aug. 4.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

106

Nutrient Requirements of Swine

Heady, E. O., R. Woodworth, D. R. Catron, and G. C. Ashton. 1954. New procedures in estimating feed substitution rates and in determining economic efficiency in pork production. Agric. Exp. Sta. Res. Bull., pp. 893–976. Iowa State College, Ames. Henry, Y., and J. Y. Dourmad. 1992. Protein nutrition and N pollution. Feed Mix. (May), pp. 25–28. Hill, G. M., G. L. Cromwell, T. D. Crenshaw, R. C. Ewan, D. A. Knabe, A. J. Lewis, D. C. Mahan, G. C. Shurson, L. L. Southern, and T. L. Veum. 1996. Impact of pharmacological intakes of zinc and (or) copper on performance of weanling pigs. J. Anim. Sci. 74(Suppl. 1):181 (Abstr.). Jongbloed, A. W., Z. Mroz, and P. A. Kemme. 1992. The effect of supplementary Aspergillus niger phytase in diets for pigs on concentration and apparent digestibility of dry matter, total phosphorus, and phytic acid in different sections of the alimentary tract. J. Anim. Sci. 70:1159–1168. ¨ Kemme, P. A., A. W. Jongbloed, Z. Mroz, and M. Makinen. 1995. Apparent ileal amino acid digestibility in pigs as affected by phytate, microbial phytase, and lactic acid. J. Anim. Sci. 73(Suppl. 1):173 (Abstr.). Kerr, B. J., and R. A. Easter. 1995. Effect of feeding reduced protein, amino acid-supplemented diets on nitrogen and energy balance in grower pigs. J. Anim. Sci. 73:3000–3008. Kornegay, E. T. 1978a. Feeding value and digestibility of soybean hulls for swine. J. Anim. Sci. 47:1272–1280. Kornegay, E. T. 1978b. Protein digestibility of hydrolyzed hog hair meal for swine. Anim. Feed Sci. Technol. 3:323–328. Kornegay, E. T. 1996. Nutritional, environmental and economical considerations for using phytase in pig diets. Pp. 279–304 in Nutrient Management of Food Animals to Enhance and Protect the Environment, E.T. Kornegay, ed. Boca Raton, FL: CRC Press Inc. Kornegay, E. T., and A. F. Harper. 1997. Environmental nutrition: Nutrient management strategies to reduce nutrient excretion of swine. The Professional Animal Scientist 13:99–111. LeMieux, F. M., L. V. Ellison, T. L. Ward, L. L. Southern, and T. D. Bidner. 1995. Excess dietary zinc for pigs weaned at 28 days. J. Anim. Sci. 74(Suppl. 1):72 (Abstr.).

Lei, X. G., P. . Ku, E. R. Miller, D. E. Ullrey, and M. T. Yokoyama. 1993. Supplemental microbial phytase improves bioavailability of dietary zinc to weanling pigs. J. Nutr. 123:1117–1123. Lei, X. G., P. K. Ku, E. R. Miller, M. T. Yokoyama, and D. E. Ullrey. 1994. Calcium level affects the efficacy of supplemental microbial phytase in corn–soybean meal diets of weanling pigs. J. Anim. Sci. 72:139–143. Mroz, Z., A. W. Jongbloed, and P. A. Kemme. 1994. Apparent digestibility and retention of nutrients bound to phytate complexes as influenced by microbial phytase and feeding regimen in pigs. J. Anim. Sci. 72:126–132. Mueller, J. P., J. P. Zublena, M. H. Poore, J. C. Barker, and J. T. Green. 1994. Managing pasture and hay fields receiving nutrients for anaerobic swine waste lagoons. N.C. Cooperative Ext. Service, AG-506. National Research Council. 1988. Nutrient Requirements of Swine. National Research Council, 9th revised ed. Washington, DC: National Academy Press. 93 pp. Pallauf, J., G. Rimbach, S. Pippig, B. Schindler, and E. Most. 1994. Effect of phytase supplementation to a phytate-rich diet based on wheat, barley and soya on the bioavailability of dietary phosphorus, calcium, magnesium, zinc and protein in piglets. Agribiol. Res. 47:39–48. Pierzynski, G. M., J. T. Sims, and G. F. Vance. 1994. Soils and Environmental Quality. Boca Raton, FL: Lewis Publishers, CRC Press. 313 pp. Radcliffe, J. S., E. T. Kornegay, and D. E. Conner, Jr. 1995. The effect of phytase on calcium release in weanling pigs fed corn–soybean meal diets. J. Anim. Sci. 73(Suppl. 1):173 (Abstr.). Sharpley, A. N., S. C. Chapra, R. Wedepohl, J. T. Sims, T. C. Daniel, and K. R. Reddy. 1994. Managing agricultural phosphorus for protection of surface waters: Issues and options. J. Environ. Qual. 23:437–451. Smith, II, J. W., M. D. Tokach, R. D. Goodband, J. L. Nelssen, W. B. Nessmith, Jr., K. Q. Owen, and B. T. Richert. 1995. The effect of increasing zinc oxide supplementation on starter pig growth performance. J. Anim. Sci. 73(Suppl. 1):72 (Abstr.). Spears, J. W. 1996. Optimizing mineral levels and sources for farm animals. Pp. 259–276 in Nutrient Management of Food Animals to Enhance and Protect the Environment, E.T. Kornegay, ed. Boca Raton, FL: CRC Press, Inc.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

9 Formulation of swine diets requires an understanding of the nutrient requirements and of the feed ingredients that can supply those nutrients. Tables 10-1 through 10-13 give summaries of nutrient requirements of various classes, weights, and levels of performance of pigs. Tables 11-1 through 11-11 give the composition of various feed ingredients and their relative values as nutrient sources. These guides can be used to formulate nutritionally adequate diets that, when fed at the recommended level, will allow pigs to perform optimally. From a nutritional standpoint, there is no ‘‘best’’ formula in terms of the ingredients that are used in the diet. Therefore, ingredients should be selected on the basis of availability, price, and quantity and quality of the nutrients that they contain. Corn, grain sorghum, barley, and wheat are the primary energy-supplying ingredients in diets for swine weighing 10 kg or more. These cereal grains are severely deficient in several essential amino acids, minerals, and vitamins. Soybean meal, other oilseed meals, and animalprotein meals are generally added as sources of supplemental amino acids to the grain, but they too are deficient in many of the essential minerals and vitamins. Table 9-1 compares the nutrient content of corn and of an unsupplemented corn–soybean meal diet with the nutrient requirements of a 40-kg growing pig. Swine diets can be formulated using rather simple mathematical procedures with a hand-held or desk calculator when a few ingredients are used in the diet. However, more sophisticated formulation procedures are needed to more precisely meet the dietary requirements on a bioavailable nutrient basis and when using larger numbers of ingredients that differ in their nutrient bioavailability. These formulation procedures often require computer programs and the expertise of a professional nutritionist. The nutrient requirements generated by the models and the feedstuff composition tables in this publication allow the user to formulate diets on the basis of bioavailable (true

Diet Formulation

or apparent ileal digestible) amino acids and bioavailable phosphorus. The procedures used to formulate diets on a bioavailable nutrient basis, though more complex, are similar to those used to formulate diets on a total nutrient basis. The following section gives examples of the calculation procedures. For the sake of simplicity, the procedures address formulation of a diet on a total nutrient basis, using corn and soybean meal as the primary feed ingredients.

FORMULATING A CORN – SOYBEAN MEAL DIET Diets can be formulated on a total nutrient basis or on an available nutrient basis. For the example given below, the formulation is on a total nutrient basis. In swine diets formulated with corn and soybean meal, the two ingredients contribute about 97.5 percent of the total diet. The remaining 2.5 percent consists of mineral supplements and carrier mixes containing vitamins, trace minerals, and additives. Corn and soybean meal are each similarly high in digestible energy (DE) concentration. Any combination of these two ingredients will result in a relatively high-energy diet. Formulation The first step in diet formulation is presented in Equation 9-1, where C is the percentage of corn and S is the percentage of dehulled soybean meal in the diet. C ` S 4 97.5

(9-1a)

S 4 97.5 1 C

(9-1b)

or

Lysine is the first limiting amino acid in corn–soybean meal diets. Because of this, one can manipulate the propor-

107

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

108

Nutrient Requirements of Swine

TABLE 9-1 Nutrients in Corn and Corn ` Soybean Meal (Dehulled) Compared with the Nutrient Requirements of a 40-kg Growing Pig of High-Medium Lean Growth Rate (325 g of carcass fat-free lean/day)

Nutrient

Corn ` Soybean Meal (74.1%:23.4%)

Corn

Indispensable amino acids (%) Arginine Histidine Isoleucine Leucine Lysine Methionine ` cystine Phenylalanine ` tyrosine Threonine Tryptophan Valine

Requirement (40-kg pig)

0.37 0.23 0.28 0.99 0.26 0.36 0.64 0.29 0.06 0.39

1.09 0.47 0.71 1.59 0.90 0.60 1.46 0.65 0.20 0.82

0.35 0.29 0.49 0.86 0.90 0.52 0.83 0.59 0.16 0.62

Mineral elements Calcium (%) Phosphorus, total (%) Phosphorus, available (%) Sodium (%) Chlorine (%) Magnesium (%) Potassium (%) Sulfur (%) Copper (mg/kg) Iodine (mg/kg) Iron (mg/kg) Manganese (mg/kg) Selenium (mg/kg) Zinc (mg/kg)

0.03 0.28 0.04 0.02 0.05 0.12 0.33 0.13 3.0 0.03 29 7.0 0.07 18

0.10 0.37 0.07 0.02 0.05 0.16 0.75 0.20 6.9 0.04 63 13.6 0.12 26

0.60 0.50 0.23 0.10 0.08 0.04 0.23 —a 4.0 0.14 60 2.0 0.15 60

Vitamins Vitamin A (IU/kg) Vitamin D (IU/kg) Vitamin E (IU/kg) Vitamin K (mg/kg) Biotin (mg/kg) Choline (g/kg) Folacin (mg/kg) Niacin, available (mg/kg) Pantothenic acid (mg/kg) Riboflavin (mg/kg) Thiamin (mg/kg) Vitamin B6 (mg/kg) Vitamin B12 (mg/kg) Ascorbic acid Linoleic acid (%)

213 0 8.3 0 0.06 0.62 0.15 0c 6.0 1.2 3.5 5.0 0 0 1.9

170 0 6.7 0 0.11 1.09 0.43 5.2 8.0 1.6 3.3 5.2 0 0 1.6

1,300 150 11 0.50b 0.05 0.30 0.30 10.0 8.0 2.5 1.0 1.0 10.0 —d 0.1

a

The requirement is unknown but is met by the sulfur from methionine and cystine. The requirement is generally met by microbial synthesis. c The niacin in cereal grain is unavailable. d The requirement is met by metabolic synthesis. b

tions of corn and dehulled soybean meal to meet the required concentration of this amino acid and be reasonably sure that the requirements for all the other essential amino acids will be met and that the amount of nonessential amino acid nitrogen will be adequate. To formulate a corn– soybean meal diet for a 40-kg pig, one may use the equation: (A 2 C)` (B 2 (97.5 1 C)) 4 (L 2 100)

(9-2)

where A is the percentage of lysine in corn, C is the percentage of corn in the diet, B is the percentage of lysine in

soybean meal, 97.5 1 C is the percentage of soybean meal in the diet, and L is the lysine requirement of the 40-kg pig, expressed as a percentage of the diet. Values for A, B, and L are then inserted into Equation 9-2, leaving only one unknown (C). The percentages of corn and soybean meal in the diet can then be solved as follows: 0.26C ` 3.02(97.5 1 C) 4 (0.90 2 100), where C is 74.1 percent corn in the diet. Because S is 97.5 1 C, then S is 23.41 percent soybean meal in the diet. The next step is to add an ingredient to supply inorganic phosphorus to complete the requirement (0.50%) for total phosphorus. If dicalcium phosphate, which contains 18.5

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Diet Formulation percent phosphorus, is selected, Equation 9-3 will show the percentage of dicalcium phosphate (DP) to include in the diet. (18.5 2 DP) 4 (0.50 2 100) 1 (74.1 2 % P in corn) 1 (23.4 2 % P in soybean meal). (18.5 2 DP) 4 (0.50 2 100) 1 (74.1 2 0.28) (9-3) 1 (23.4 2 0.69). DP 4 0.71% dicalcium phosphate in diet. The next step is to add an ingredient to supply calcium to complete the requirement for calcium (0.60%). If ground limestone, which contains 38 percent calcium, is selected, Equation 9-4 will show the percentage of ground limestone (GL) to include in the diet. (38 2 GL) 4 (0.60 2 100) 1 (74.1 2 % Ca in corn) 1 (23.4 2 % Ca in soybean meal) 1 (0.71 2 % Ca in dicalcium phosphate) (38 2 GL) 4 (0.60 2 100) (9-4) 1 (74.1 2 0.03) 1 (23.4 2 0.34) 1 (0.71 2 22). GL 4 0.90% ground limestone in diet One can completely fortify the swine diet by adding 0.25 percent sodium chloride; a vitamin premix that supplies the vitamins deficient in the corn–soybean meal mixture (vitamins A, D, E, K, B12, riboflavin, niacin, pantothenic acid); a trace mineral premix that supplies the trace minerals that may be deficient (iron, zinc, copper, manganese, iodine, and selenium); and if desired, a premix that contains one or more antimicrobial agents. The fortified diet is

TABLE 9-2

109

Fortified Swine Diet

Nutrient Corn Soybean meal, dehulled Dicalcium phosphate Ground limestone Sodium chloride Vitamin premix Trace mineral premix Antimicrobial premix Total

Percent 74.44 23.40 0.71 0.90 0.25 0.10 0.10 0.10 100.00

shown in Table 9-2. The diet is made to total 100 percent by increasing the amount of corn to 74.44 percent. Formulation on a true or apparent digestible lysine basis is essentially the same as described above except that the true or apparent digestible lysine values for corn and soybean meal are used in the calculations. These values are obtained by multiplying the total lysine concentrations in Table 11-4 by the amino acid digestibility coefficients in Tables 11-5 or 11-6. For example, to meet the apparent digestible lysine requirement (0.73%) of the same 40-kg pig in the previous example, the apparent digestible lysine in corn (0.26% total lysine 2 66% apparent digestibility of lysine 4 0.17% apparent digestible lysine) and soybean meal (3.02% 2 85% 4 2.56%) are used. Similar procedures are used to formulate diets on an available phosphorus basis. Based on the composition and bioavailability data in Table 11-1, the bioavailable phosphorus requirement (0.23%) of the 40-kg pig is met by using the bioavailable phosphorus in corn (0.28% total phosphorus 2 14% bioavailability 4 0.039% bioavailable phosphorus), soybean meal (0.69% 2 23% 4 0.16%), and dicalcium phosphate (18.5% 2 100% 4 18.5%).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

10 Nutrient requirements for starting, growing, and finishing pigs, gestating and lactating sows, and sexually active boars are provided in the tables of this chapter. Estimates are listed for energy, amino acids, minerals, vitamins, and linoleic acid. The amino acid requirements are expressed on a true and apparent ileal digestible basis, which applies to all types of feed ingredients. They are also expressed on a total basis, which applies to a corn–soybean mealbased diet. For minerals and vitamins, the requirements include the amounts of these nutrients that are provided by feed ingredients. Thus they are not estimates of nutrient quantities that should be added to diets. Tables 10-1 to 10-6 give estimated requirements of young weanling pigs from 3 to 20 kg, and of growingfinishing pigs from 20 to 120 kg body weight. The amino acid requirements in Table 10-1 are generated by the model described in Chapter 3 for pigs (equal ratio of barrows and gilts) of a high-medium lean growth rate (325 g of carcass fat-free gain/day) from 20 to 120 kg and housed under ideal temperature and space conditions. Table 10-3 gives separate requirements for barrows and gilts of three lean growth rates from 50 to 120 kg. Tables 10-2 and 10-4 give estimated daily requirements of amino acids. Requirements for minerals, vitamins, and linoleic acid are given both on a dietary concentration (Table 10-5) and daily intake (Table 10-6) basis. Amino acid requirements, estimated by the sow models, for gestating sows of various breeding weights, gestation weight gains, and anticipated litter sizes and for lactating sows of various postfarrowing weights, lactation weight changes, and weight gains of their pigs are given in Tables 10-7 through 10-10. The estimates are based on ideal temperature environments. Dietary concentrations and daily intake requirements of minerals, vitamins, and linoleic acid are given in Tables 10-11 and 10-12, respectively. Table 10-13 lists estimated requirements of sexually active boars.

Nutrient Requirement Tables

The amino acid requirements in the tables are given as examples. The models included in this publication allow the user to generate tables of estimated amino acid requirements for swine under various conditions (i.e., different lean growth rates, feed intakes, energy density of diets, environmental temperature, floor space, etc.). The mineral and vitamin estimates in these tables, however, are the committee’s best estimates of the dietary requirements for average pigs under average conditions. The growth model may generate slightly different estimates of mineral and vitamin requirements because it uses an exponential equation to estimate the requirements at various body weights of growing pigs (3 to 120 kg), based on the coefficients given in Table 3-2. The requirements for certain minerals and/or vitamins by pigs possessing a high lean growth rate, due to superior genetics or high health status, may be higher than the levels shown in the tables, but definitive information was not available to estimate a higher quantitative requirement. Slightly higher levels of calcium and phosphorus than shown in the tables are required by developing boars and replacement gilts from 50 to 120 kg body weight, and suggestions are appropriately footnoted in Tables 10-5 and 10-6. The requirements listed in the following tables do not include any intentional surpluses. They are the committee’s best estimates of minimum requirements. In practice, however, a margin of safety is commonly added to the stated requirements, and these levels are often referred to as nutrient ‘‘allowances.’’ Nutrient allowances are generally established by professional nutritionists to account for variability in nutrient composition and in nutrient bioavailability of feedstuffs, presence of inhibitors or toxins in ingredients, inadequate processing or mixing of diets, partial loss of nutrients from storage, and other factors. Because of these factors, the statement on a feed label that the product ‘‘meets or exceeds National Research

110

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables Council requirements’’ by itself should not be accepted as prima facie evidence of a complete and balanced diet. Knowledge of the nutritional constraints and limitations is

TABLE 10-1

111

important for the proper use of the requirement tables that follow.

Dietary Amino Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90% dry matter)a Body Weight (kg)

Average weight in range (kg) DE content of diet (kcal/kg) ME content of diet (kcal/kg)b Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)b Estimated feed intake (g/day) Crude protein (%)c

3–5

5–10

10–20

20–50

50–80

80–120

4 3,400 3,265 855 820 250 26.0

7.5 3,400 3,265 1,690 1,620 500 23.7

15 3,400 3,265 3,400 3,265 1,000 20.9

35 3,400 3,265 6,305 6,050 1,855 18.0

65 3,400 3,265 8,760 8,410 2,575 15.5

100 3,400 3,265 10,450 10,030 3,075 13.2

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.54 0.43 0.73 1.35 1.34 0.36 0.76 0.80 1.26 0.84 0.24 0.91

0.49 0.38 0.65 1.20 1.19 0.32 0.68 0.71 1.12 0.74 0.22 0.81

Amino acid requirementsd True ileal digestible basis (%) 0.42 0.33 0.32 0.26 0.55 0.45 1.02 0.83 1.01 0.83 0.27 0.22 0.58 0.47 0.61 0.49 0.95 0.78 0.63 0.52 0.18 0.15 0.69 0.56

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.51 0.40 0.69 1.29 1.26 0.34 0.71 0.75 1.18 0.75 0.22 0.84

0.46 0.36 0.61 1.15 1.11 0.30 0.63 0.66 1.05 0.66 0.19 0.74

Apparent ileal digestible basis (%) 0.39 0.31 0.31 0.25 0.52 0.42 0.98 0.80 0.94 0.77 0.26 0.21 0.53 0.44 0.56 0.46 0.89 0.72 0.56 0.46 0.16 0.13 0.63 0.51

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.59 0.48 0.83 1.50 1.50 0.40 0.86 0.90 1.41 0.98 0.27 1.04

0.54 0.43 0.73 1.32 1.35 0.35 0.76 0.80 1.25 0.86 0.24 0.92

0.46 0.36 0.63 1.12 1.15 0.30 0.65 0.68 1.06 0.74 0.21 0.79

Total basis (%)e 0.37 0.30 0.51 0.90 0.95 0.25 0.54 0.55 0.87 0.61 0.17 0.64

a

0.24 0.21 0.37 0.67 0.66 0.18 0.39 0.40 0.63 0.43 0.12 0.45

0.16 0.16 0.29 0.51 0.52 0.14 0.31 0.31 0.49 0.34 0.10 0.35

0.22 0.20 0.34 0.64 0.61 0.17 0.36 0.37 0.58 0.37 0.10 0.41

0.14 0.16 0.26 0.50 0.47 0.13 0.29 0.28 0.45 0.30 0.08 0.32

0.27 0.24 0.42 0.71 0.75 0.20 0.44 0.44 0.70 0.51 0.14 0.52

0.19 0.19 0.33 0.54 0.60 0.16 0.35 0.34 0.55 0.41 0.11 0.40

Mixed gender (1:1 ratio of barrows to gilts) of pigs with high-medium lean growth rate (325 g/day of carcass fat-free lean) from 20 to 120 kg body weight. Assumes that ME is 96% of DE. In corn–soybean meal diets of these crude protein levels, ME is 94–96% of DE. Crude protein levels apply to corn–soybean meal diets. In 3–10 kg pigs fed diets with dried plasma and/or dried milk products, protein levels will be 2–3% less than shown. d Total amino acid requirements are based on the following types of diets: 3–5 kg pigs, corn–soybean meal diet that includes 5% dried plasma and 25–50% dried milk products; 5–10 kg pigs, corn–soybean meal diet that includes 5 to 25% dried milk products; 10–120 kg pigs, corn–soybean meal diet. e The total lysine percentages for 3–20 kg pigs are estimated from empirical data. The other amino acids for 3–20 kg pigs are based on the ratios of amino acids to lysine (true digestible basis); however, there are very few empirical data to support these ratios. The requirements for 20–120 kg pigs are estimated from the growth model. b c

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

112

Nutrient Requirements of Swine

TABLE 10-2

Daily Amino Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90% dry matter)a

Average weight in range (kg) DE content of diet (kcal/kg) ME content of diet (kcal/kg)b Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)b Estimated feed intake (g/day) Crude protein (%)c

Body Weight (kg) 3–5 5–10

10–20

20–50

50–80

80–120

4 3,400 3,265 855 820 250 26.0

15 3,400 3,265 3,400 3,265 1,000 20.9

35 3,400 3,265 6,305 6,050 1,855 18.0

65 3,400 3,265 8,760 8,410 2,575 15.5

100 3,400 3,265 10,450 10,030 3,075 13.2

6.2 5.5 9.4 7.2 17.1 4.6 10.0 10.2 16.1 11.0 3.1 11.6

4.8 5.1 8.8 15.8 15.8 4.3 9.5 9.4 15.1 10.5 2.9 10.8

7.5 3,400 3,265 1,690 1,620 500 23.7

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

1.4 1.1 1.8 3.4 3.4 0.9 1.9 2.0 3.2 2.1 0.6 2.3

2.4 1.9 3.2 6.0 5.9 1.6 3.4 3.5 5.5 3.7 1.1 4.0

Amino acid requirementsd True ileal digestible basis (g/day) 4.2 6.1 3.2 4.9 5.5 8.4 10.3 15.5 10.1 15.3 2.7 4.1 5.8 8.8 6.1 9.1 9.5 14.4 6.3 9.7 1.9 2.8 6.9 10.4

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

1.3 1.0 1.7 3.2 3.2 0.9 1.8 1.9 3.0 1.9 0.5 2.1

2.3 1.8 3.0 5.7 5.5 1.5 3.1 3.3 5.2 3.3 1.0 3.7

Apparent ileal digestible basis (g/day) 3.9 5.7 3.1 4.6 5.2 7.8 9.8 14.8 9.4 14.2 2.6 3.9 5.3 8.2 5.7 8.5 8.9 13.4 5.6 8.5 1.6 2.4 6.3 9.5

5.7 5.2 8.7 16.5 15.8 4.4 9.3 9.4 15.0 9.6 2.7 10.6

4.3 4.8 8.0 15.3 14.4 4.1 8.8 8.6 13.9 9.1 2.5 9.8

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

1.5 1.2 2.1 3.8 3.8 1.0 2.2 2.3 3.5 2.5 0.7 2.6

2.7 2.1 3.7 6.6 6.7 1.8 3.8 4.0 6.2 4.3 1.2 4.6

Total basis (g/day)e 6.8 5.6 9.5 16.8 17.5 4.6 9.9 10.2 16.1 11.3 3.2 11.9

7.1 6.3 10.7 18.4 19.7 5.1 11.3 11.3 18.0 13.0 3.6 13.3

5.7 5.9 10.1 16.6 18.5 4.8 10.8 10.4 16.8 12.6 3.4 12.4

4.6 3.7 6.3 11.2 11.5 3.0 6.5 6.8 10.6 7.4 2.1 7.9

a

Mixed gender (1:1 ratio of barrows to gilts) of pigs with high-medium lean growth rate (325 g/day of carcass fat-free lean) from 20 to 120 kg body weight. Assumes that ME is 96% of DE. In corn–soybean meal diets of these crude protein levels, ME is 94–96% of DE. c Crude protein levels apply to corn–soybean meal diets. In 3–10 kg pigs fed diets with dried plasma and/or dried milk products, protein levels will be 2–3% less than shown. d Total amino acid requirements are based on the following types of diets: 3–5 kg pigs, corn–soybean meal diet that includes 5% dried plasma and 25–50% dried milk products; 5–10 kg pigs, corn–soybean meal diet that includes 5 to 25% dried milk products; 10–120 kg pigs, corn–soybean meal diet. e The total lysine estimates for 3–20 kg pigs are calculated by multiplying the percentages in Table 10–1 (estimated from empirical data) by the estimated feed intake. The other amino acids for 3–20 kg pigs are based on the ratios of amino acids to lysine (true digestible basis); however, there are very few empirical data to support these ratios. The estimates for 20–120 kg pigs are from the growth model. b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables

113

TABLE 10-3 Dietary Amino Acid Requirements of Barrows and Gilts of Different Lean Growth Rates and Allowed Feed Ad Libitum (90% dry matter)a Body weight range:

50–80 kg Body Weight

Lean gain (g/day) Gender

300 Barrow

300 Gilt

325 Barrow

325 Gilt

350 Barrow

350 Gilt

300 Barrow

80–120 kg Body Weight

Average weight in range (kg) DE content of diet (kcal/kg) ME content of diet (kcal/kg)b Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)b Estimated feed intake (g/day) Crude protein (%)c

65 3,400 3,265 9,360 8,985 2,750 14.2

65 3,400 3,265 8,165 7,840 2,400 15.5

65 3,400 3,265 9,360 8,985 2,755 14.9

65 3,400 3,265 8,165 7,840 2,400 16.3

65 3,400 3,265 9,360 8,985 2,755 15.6

65 3,400 3,265 8,165 7,840 2,400 17.1

100 3,400 3,265 11,150 10,705 3,280 12.2

300 Gilt

325 Barrow

100 3,400 3,265 9,750 9,360 2,865 13.2

100 3,400 3,265 11,150 10,705 3,280 12.7

325 Gilt

350 Barrow

100 3,400 3,265 9,750 9,360 2,865 13.8

100 3,400 3,265 11,150 10,705 3,280 13.2

350 Gilt 100 3,400 3,265 9,750 9,360 2,865 14.4

Amino acid requirementsd True ileal digestible basis (%) Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.20 0.18 0.32 0.58 0.58 0.16 0.34 0.34 0.54 0.37 0.11 0.39

0.23 0.21 0.36 0.66 0.66 0.18 0.39 0.39 0.62 0.43 0.12 0.45

0.22 0.20 0.34 0.62 0.62 0.17 0.36 0.37 0.59 0.40 0.11 0.42

0.26 0.23 0.39 0.72 0.71 0.19 0.42 0.42 0.67 0.46 0.13 0.48

0.25 0.21 0.37 0.67 0.67 0.18 0.39 0.40 0.63 0.43 0.12 0.45

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.19 0.17 0.29 0.56 0.53 0.15 0.31 0.32 0.50 0.32 0.09 0.36

0.21 0.20 0.34 0.64 0.61 0.17 0.36 0.36 0.58 0.37 0.10 0.41

0.21 0.19 0.31 0.60 0.57 0.16 0.34 0.34 0.54 0.35 0.10 0.38

0.24 0.21 0.36 0.69 0.66 0.18 0.39 0.39 0.62 0.40 0.11 0.44

0.23 0.20 0.34 0.65 0.61 0.17 0.36 0.37 0.58 0.37 0.10 0.41

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.24 0.21 0.36 0.61 0.67 0.17 0.38 0.38 0.61 0.44 0.12 0.45

0.27 0.24 0.41 0.71 0.76 0.20 0.44 0.44 0.70 0.50 0.14 0.51

0.26 0.23 0.39 0.67 0.72 0.19 0.41 0.41 0.65 0.47 0.13 0.48

0.29 0.26 0.45 0.77 0.82 0.21 0.47 0.47 0.75 0.54 0.15 0.55

0.28 0.24 0.42 0.72 0.77 0.20 0.44 0.44 0.70 0.51 0.14 0.52

0.28 0.24 0.42 0.77 0.76 0.21 0.44 0.46 0.72 0.49 0.14 0.52

0.13 0.14 0.25 0.45 0.45 0.12 0.27 0.27 0.43 0.30 0.08 0.30

0.15 0.16 0.29 0.51 0.51 0.14 0.31 0.30 0.49 0.34 0.10 0.35

0.15 0.15 0.27 0.48 0.48 0.13 0.29 0.29 0.46 0.32 0.09 0.33

0.17 0.18 0.31 0.55 0.55 0.15 0.33 0.33 0.52 0.37 0.10 0.38

0.16 0.17 0.29 0.52 0.52 0.14 0.31 0.31 0.49 0.34 0.10 0.35

0.19 0.19 0.33 0.59 0.59 0.16 0.35 0.35 0.56 0.39 0.11 0.40

0.13 0.15 0.26 0.50 0.47 0.13 0.29 0.28 0.45 0.30 0.08 0.32

0.13 0.15 0.24 0.47 0.44 0.13 0.27 0.26 0.42 0.28 0.07 0.30

0.15 0.17 0.28 0.53 0.51 0.14 0.31 0.30 0.49 0.32 0.09 0.34

0.15 0.16 0.26 0.50 0.47 0.13 0.29 0.28 0.45 0.30 0.08 0.32

0.17 0.18 0.30 0.57 0.54 0.15 0.33 0.32 0.52 0.34 0.09 0.37

0.18 0.19 0.33 0.54 0.60 0.15 0.35 0.34 0.54 0.41 0.11 0.40

0.18 0.18 0.31 0.50 0.57 0.15 0.33 0.32 0.51 0.38 0.10 0.38

0.20 0.20 0.35 0.58 0.64 0.17 0.38 0.36 0.59 0.44 0.12 0.43

0.19 0.19 0.33 0.54 0.60 0.16 0.35 0.34 0.55 0.41 0.11 0.40

0.22 0.22 0.37 0.63 0.69 0.18 0.40 0.39 0.63 0.46 0.13 0.46

Apparent ileal digestible basis (%) 0.26 0.23 0.39 0.74 0.71 0.20 0.41 0.42 0.67 0.43 0.12 0.47

0.12 0.14 0.23 0.43 0.41 0.12 0.25 0.24 0.39 0.26 0.07 0.28

Total basis (%)c 0.32 0.28 0.48 0.83 0.88 0.23 0.50 0.51 0.80 0.58 0.16 0.59

0.16 0.17 0.29 0.46 0.53 0.14 0.31 0.29 0.48 0.36 0.10 0.35

a Average lean growth rates of 300, 325, and 350 g/day of carcass fat-free lean represent pigs with medium, high-medium, and high lean growth rates from 20 to 120 kg body weight. b Assumes that ME is 96% of DE. c Crude protein and total amino acid requirements are based on a corn–soybean meal diet. d Estimated from the growth model.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

114

Nutrient Requirements of Swine

TABLE 10-4 Daily Amino Acid Requirements of Barrows and Gilts of Different Lean Growth Rates and Allowed Feed Ad Libitum (90% dry matter)a Body weight range:

50–80 kg Body Weight

Lean gain (g/day) Gender

300 Barrow

300 Gilt

325 Barrow

325 Gilt

350 Barrow

350 Gilt

300 Barrow

80–120 kg Body Weight

Average weight in range (kg) DE content of diet (kcal/kg) ME content of diet (kcal/kg)b Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)b Estimated feed intake (g/day) Crude protein (%)c

65 3,400 3,265 9,360 8,985 2,755 14.2

65 3,400 3,265 8,165 7,840 2,400 15.5

65 3,400 3,265 9,360 8,985 2,755 14.9

65 3,400 3,265 8,165 7,840 2,400 16.3

65 3,400 3,265 9,360 8,985 2,755 15.6

65 3,400 3,265 8,165 7,840 2,400 17.1

100 3,400 3,265 11,150 10,705 3,280 12.2

300 Gilt 100 3,400 3,265 9,750 9,360 2,865 13.2

325 Barrow 100 3,400 3,265 11,150 10,705 3,280 12.7

325 Gilt 100 3,400 3,265 9,750 9,360 2,865 13.8

350 Barrow 100 3,400 3,265 11,150 10,705 3,280 13.2

350 Gilt 100 3,400 3,265 9,750 9,360 2,865 14.4

Amino acid requirementsd True ileal digestible basis (g/day) Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

5.6 5.1 8.7 15.9 15.9 4.3 9.3 9.4 15.0 10.3 2.9 10.8

6.2 5.5 9.4 17.2 17.1 4.6 10.0 10.2 16.1 11.0 3.1 11.6

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

5.1 4.8 8.0 15.3 14.6 4.1 8.6 8.7 13.9 8.9 2.5 9.8

5.7 5.2 8.7 16.5 15.7 4.4 9.3 9.4 15.0 9.6 2.7 10.6

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

6.4 5.8 10.0 16.9 18.3 4.8 10.5 10.5 16.7 12.2 3.3 12.4

7.1 6.3 10.7 18.4 19.7 5.1 11.3 11.3 18.0 13.0 3.6 13.3

6.8 5.9 10.1 18.5 18.4 5.0 10.7 10.9 17.3 11.8 3.4 12.5

4.2 4.7 8.2 14.6 14.7 4.0 8.9 8.7 14.0 9.9 2.7 10.0

4.8 5.1 8.8 15.8 15.8 4.3 9.5 9.4 15.1 10.5 2.9 10.8

5.3 5.4 9.4 16.9 17.0 4.6 10.1 10.1 16.1 11.2 3.2 11.5

3.8 4.4 7.5 14.2 13.4 3.8 8.3 8.0 12.9 8.5 2.3 9.1

4.3 4.8 8.0 15.3 14.4 4.1 8.8 8.6 13.9 9.1 2.5 9.8

4.8 5.1 8.6 16.4 15.5 4.4 9.4 9.3 14.9 9.7 2.6 10.5

5.1 5.5 9.4 15.3 17.3 4.4 10.1 9.7 15.6 11.8 3.2 11.5

5.7 5.9 10.1 16.6 18.5 4.8 10.8 10.4 16.8 12.6 3.4 12.4

6.3 6.3 10.7 17.9 19.7 5.1 11.5 11.2 18.0 13.3 3.6 13.2

Apparent ileal digestible basis (g/day) 6.3 5.5 9.3 17.7 16.9 4.7 9.9 10.1 16.1 10.3 2.9 11.4

Total basis (g/day)c 7.7 6.7 11.5 19.8 21.1 5.5 12.1 12.2 19.3 13.9 3.8 14.3

a Average lean growth rates of 300, 325, and 350 g/day of carcass fat-free lean represent pigs with medium, high-medium, and high lean growth rates from 20 to 120 kg body weight. b Assumes that ME is 96% of DE. c Crude protein and total amino acid requirements are based on a corn–soybean meal diet. d Estimated from the growth model.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables

115

TABLE 10-5 Dietary Mineral, Vitamin, and Fatty Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90% dry matter) a Body Weight (kg) Average weight in range (kg) DE content of diet (kcal/kg) ME content of diet (kcal/kg)b Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)b Estimated feed intake (g/day)

3–5

5–10

10–20

20–50

50–80

80–120

4 3,400 3,265 855 820 250

7.5 3,400 3,265 1,690 1,620 500

15 3,400 3,265 3,400 3,265 1,000

35 3,400 3,265 6,305 6,050 1,855

65 3,400 3,265 8,760 8,410 2,575

100 3,400 3,265 10,450 10,030 3,075

Requirements (% or amount/kg of diet) Mineral elements Calcium (%)c Phosphorus, total (%)c Phosphorus, available (%)c Sodium (%) Chlorine (%) Magnesium (%) Potassium (%) Copper (mg) Iodine (mg) Iron (mg) Manganese (mg) Selenium (mg) Zinc (mg) Vitamins Vitamin A (IU)d Vitamin D3 (IU)d Vitamin E (IU)d Vitamin K (menadione) (mg) Biotin (mg) Choline (g) Folacin (mg) Niacin, available (mg) e Pantothenic acid (mg) Riboflavin (mg) Thiamin (mg) Vitamin B6 (mg) Vitamin B12 (mg) Linoleic acid (%)

0.90 0.70 0.55 0.25 0.25 0.04 0.30 6.00 0.14 100 4.00 0.30 100

0.80 0.65 0.40 0.20 0.20 0.04 0.28 6.00 0.14 100 4.00 0.30 100

0.70 0.60 0.32 0.15 0.15 0.04 0.26 5.00 0.14 80 3.00 0.25 80

0.60 0.50 0.23 0.10 0.08 0.04 0.23 4.00 0.14 60 2.00 0.15 60

0.50 0.45 0.19 0.10 0.08 0.04 0.19 3.50 0.14 50 2.00 0.15 50

0.45 0.40 0.15 0.10 0.08 0.04 0.17 3.00 0.14 40 2.00 0.15 50

2,200 220 16 0.50 0.08 0.60 0.30 20.00 12.00 4.00 1.50 2.00 20.00

2,200 220 16 0.50 0.05 0.50 0.30 15.00 10.00 3.50 1.00 1.50 17.50

1,750 200 11 0.50 0.05 0.40 0.30 12.50 9.00 3.00 1.00 1.50 15.00

1,300 150 11 0.50 0.05 0.30 0.30 10.00 8.00 2.50 1.00 1.00 10.00

1,300 150 11 0.50 0.05 0.30 0.30 7.00 7.00 2.00 1.00 1.00 5.00

1,300 150 11 0.50 0.05 0.30 0.30 7.00 7.00 2.00 1.00 1.00 5.00

0.10

0.10

0.10

0.10

0.10

0.10

a

Pigs of mixed gender (1:1 ratio of barrows to gilts). The requirements of certain minerals and vitamins may be slightly higher for pigs having high lean growth rates (.325 g/day of carcass fat-free lean), but no distinction is made. b Assumes that ME is 96% of DE. In corn–soybean meal diets, ME is 94–96% of DE, depending on crude protein level of the diet. c The percentages of calcium, phosphorus, and available phosphorus should be increased by 0.05 to 0.1 percentage points for developing boars and replacement gilts from 50 to 120 kg body weight. d Conversions: 1 IU vitamin A 4 0.344 mg (g retinyl acetate; 1 IU vitamin D3 4 0.025 mg cholecalciferol; 1 IU vitamin E 4 0.67 mg of D-a-tocopherol or 1 mg of DLa-tocopheryl acetate. e The niacin in corn, grain sorghum, wheat, and barley is unavailable. Similarly, the niacin in by-products made from these cereal grains is poorly available unless the byproducts have undergone a fermentation or wet-milling process.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

116

Nutrient Requirements of Swine

TABLE 10-6 dry matter) a

Daily Mineral, Vitamin, and Fatty Acid Requirements of Growing Pigs Allowed Feed Ad Libitum (90% Body Weight (kg)

Average weight in range (kg) DE content of diet (kcal/kg) ME content of diet (kcal/kg) b Estimated DE intake (kcal/day) Estimated ME intake (kcal/day) b Estimated feed intake (g/day)

3–5

5–10

10–20

20–50

50–80

80–120

4 3,400 3,265 855 820 250

7.5 3,400 3,265 1,690 1,620 500

15 3,400 3,265 3,400 3,265 1,000

35 3,400 3,265 6,305 6,050 1,855

65 3,400 3,265 8,760 8,410 2,575

100 3,400 3,265 10,450 10,030 3,075

Requirements (amount/day) Mineral elements Calcium (g) c Phosphorus, total (g) c Phosphorus, available (g) c Sodium (g) Chlorine (g) Magnesium (g) Potassium (g) Copper (mg) Iodine (mg) Iron (mg) Manganese (mg) Selenium (mg) Zinc (mg) Vitamins Vitamin A (IU) d Vitamin D3 (IU) d Vitamin E (IU) d Vitamin K (menadione) (mg) Biotin (mg) Choline (g) Folacin (mg) Niacin, available (mg) e Pantothenic acid (mg) Riboflavin (mg) Thiamin (mg) Vitamin B6 (mg) Vitamin B12 (mg) Linoleic acid (g)

2.25 1.75 1.38 0.63 0.63 0.10 0.75 1.50 0.04 25.00 1.00 0.08 25.00

4.00 3.25 2.00 1.00 1.00 0.20 1.40 3.00 0.07 50.00 2.00 0.15 50.00

7.00 6.00 3.20 1.50 1.50 0.40 2.60 5.00 0.14 80.00 3.00 0.25 80.00

11.13 9.28 4.27 1.86 1.48 0.74 4.27 7.42 0.26 111.30 3.71 0.28 111.30

12.88 11.59 4.89 2.58 2.06 1.03 4.89 9.01 0.36 129.75 5.15 0.39 129.75

13.84 12.30 4.61 3.08 2.46 1.23 5.23 9.23 0.43 123.00 6.15 0.46 153.75

550 55 4 0.13 0.02 0.15 0.08 5.00 3.00 1.00 0.38 0.50 5.00

1,100 110 8 0.25 0.03 0.25 0.15 7.50 5.00 1.75 0.50 0.75 8.75

1,750 200 11 0.50 0.05 0.40 0.30 12.50 9.00 3.00 1.00 1.50 15.00

2,412 278 20 0.93 0.09 0.56 0.56 18.55 14.84 4.64 1.86 1.86 18.55

3,348 386 28 1.29 0.13 0.77 0.77 18.03 18.03 5.15 2.58 2.58 12.88

3,998 461 34 1.54 0.15 0.92 0.92 21.53 21.53 6.15 3.08 3.08 15.38

0.25

0.50

1.00

1.86

2.58

3.08

a

Pigs of mixed gender (1:1 ratio of barrows to gilts). The daily requirements of certain minerals and vitamins may be slightly higher for pigs having high lean growth rates (.325 g/day of carcass fat-free lean), but no distinction is made. b Assumes that ME is 96% of DE. In corn–soybean meal diets, ME is 94–96% of DE, depending on crude protein level of the diet. c The daily amounts of calcium, phosphorus, and available phosphorus are slightly higher in developing boars and gilts from 50 to 120 kg body weight. d Conversions: 1 IU vitamin A 4 0.344 mg (g retinyl acetate; 1 IU vitamin D3 4 0.025 mg cholecalciferol; 1 IU vitamin E 4 0.67 mg of D-a-tocopherol or 1 mg of DLa-tocopheryl acetate. e The niacin in corn, grain sorghum, wheat, and barley is unavailable. Similarly, the niacin in by-products made from these cereal grains is poorly available unless the byproducts have undergone a fermentation or wet-milling process.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables TABLE 10-7

117

Dietary Amino Acid Requirements of Gestating Sows (90% dry matter) a Body Weight at Breeding (kg) 125 150 Gestation Weight Gain (kg)b

DE content of diet (kcal/kg) ME content of diet (kcal/kg) c Estimated DE intake (kcal/day) Estimated ME intake (kcal/day) c Estimated feed intake (kg/day) Crude protein (%) d

175

200

200

200

55 45 Anticipated Pigs in Litter 11 12

40

35

30

35

12

12

12

14

3,400 3,265 6,660 6,395 1.96 12.9

3,400 3,265 6,405 6,150 1.88 12.4

3,400 3,265 6,535 6,275 1.92 12.0

3,400 3,265 6,115 5,870 1.80 12.1

3,400 3,265 6,275 6,025 1.85 12.4

0.00 0.14 0.26 0.41 0.44 0.12 0.32 0.25 0.44 0.37 0.09 0.30

0.00 0.15 0.27 0.44 0.46 0.13 0.33 0.27 0.46 0.38 0.09 0.31

0.00 0.13 0.24 0.40 0.40 0.12 0.30 0.23 0.41 0.32 0.07 0.27

0.00 0.14 0.25 0.43 0.42 0.12 0.31 0.24 0.43 0.33 0.08 0.28

0.00 0.17 0.30 0.43 0.52 0.13 0.36 0.28 0.49 0.44 0.10 0.34

0.00 0.17 0.31 0.45 0.54 0.14 0.37 0.30 0.51 0.45 0.11 0.36

3,400 3,265 6,265 6,015 1.84 12.8

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.04 0.16 0.29 0.48 0.50 0.14 0.33 0.29 0.48 0.37 0.10 0.34

0.00 0.16 0.28 0.47 0.49 0.13 0.33 0.28 0.48 0.38 0.10 0.33

Amino acid requirements True ileal digestible basis (%) 0.00 0.00 0.15 0.14 0.27 0.26 0.44 0.41 0.46 0.44 0.13 0.12 0.32 0.31 0.27 0.25 0.46 0.44 0.37 0.36 0.09 0.09 0.31 0.30

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.03 0.15 0.26 0.47 0.45 0.13 0.30 0.27 0.45 0.32 0.08 0.31

0.00 0.15 0.26 0.46 0.45 0.13 0.31 0.26 0.44 0.33 0.08 0.30

Apparent ileal digestible basis (%) 0.00 0.00 0.14 0.13 0.25 0.24 0.43 0.40 0.42 0.40 0.12 0.11 0.30 0.29 0.24 0.23 0.42 0.40 0.32 0.31 0.08 0.07 0.28 0.27

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.06 0.19 0.33 0.50 0.58 0.15 0.37 0.32 0.54 0.44 0.11 0.39

0.03 0.18 0.32 0.49 0.57 0.15 0.38 0.32 0.54 0.45 0.11 0.38

0.00 0.17 0.31 0.46 0.54 0.14 0.37 0.30 0.51 0.44 0.11 0.36

Total basis (%)d 0.00 0.16 0.30 0.42 0.52 0.13 0.36 0.28 0.49 0.43 0.10 0.34

a

Daily intakes of DE and feed and the amino acid requirements are estimated by the gestation model. Weight gain includes maternal tissue and products of conception. c Assumes that ME is 96% of DE. d Crude protein and total amino acid requirements are based on a corn–soybean meal diet. b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

118

Nutrient Requirements of Swine

TABLE 10-8

Daily Amino Acid Requirements of Gestating Sows (90% dry matter) a Body Weight at Breeding (kg) 125 150 Gestation Weight Gain (kg) b

DE content of diet (kcal/kg) ME content of diet (kcal/kg)c Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)c Estimated feed intake (kg/day) Crude protein (%)d

175

200

200

200

55 45 Anticipated Pigs in Litter 11 12

40

35

30

35

12

12

12

14

3,400 3,265 6,660 6,395 1.96 12.9

3,400 3,265 6,405 6,150 1.88 12.4

3,400 3,265 6,535 6,275 1.92 12.0

3,400 3,265 6,115 5,870 1.80 12.1

3,400 3,265 6,275 6,025 1.85 12.4

3,400 3,265 6,265 6,015 1.84 12.8

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.8 3.1 5.6 9.4 9.7 2.7 6.4 5.7 9.5 7.3 1.9 6.6

0.1 2.9 5.2 8.7 9.0 2.5 6.1 5.2 8.9 7.0 1.8 6.1

Amino acid requirements True ileal digestible basis (g/day) 0.0 0.0 2.8 2.7 5.1 5.0 8.3 7.9 8.7 8.4 2.4 2.3 6.1 6.0 5.0 4.8 8.6 8.4 6.9 6.9 1.7 1.7 5.9 5.7

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.6 2.9 5.1 9.2 8.9 2.5 6.0 5.2 8.8 6.3 1.6 6.0

0.0 2.7 4.8 8.4 8.2 2.4 5.7 4.8 8.2 6.0 1.5 5.6

Apparent ileal digestible basis (g/day) 0.0 0.0 2.6 2.5 4.7 4.5 8.1 7.7 7.9 7.6 2.3 2.2 5.7 5.6 4.6 4.4 8.0 7.7 6.0 6.0 1.4 1.4 5.4 5.2

0.0 2.4 4.3 7.3 7.2 2.1 5.3 4.2 7.3 5.7 1.3 4.9

0.0 2.6 4.6 7.9 7.7 2.2 5.7 4.5 7.9 6.1 1.4 5.3

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

1.3 3.6 6.4 9.9 11.4 2.9 7.3 6.3 10.6 8.6 2.2 7.6

0.5 3.4 6.0 9.0 10.6 2.7 7.0 5.8 9.9 8.3 2.0 7.0

Total basis (g/day) d 0.0 3.2 5.7 8.2 9.9 2.6 6.8 5.4 9.4 8.2 1.9 6.6

0.0 3.0 5.4 7.7 9.4 2.4 6.5 5.0 8.9 7.8 1.8 6.2

0.0 3.2 5.8 8.3 10.0 2.6 6.9 5.4 9.5 8.3 2.0 6.7

0.0 3.3 5.9 8.6 10.3 2.6 6.9 5.6 9.6 8.3 2.0 6.8

a

Daily intakes of DE and feed and the amino acid requirements are estimated by the gestation model. Weight gain includes maternal tissue and products of conception. c Assumes that ME is 96% of DE. d Crude protein and total amino acid requirements are based on a corn–soybean meal diet. b

Copyright © National Academy of Sciences. All rights reserved.

0.0 2.5 4.7 7.4 7.9 2.2 5.7 4.6 7.9 6.6 1.6 5.4

0.0 2.7 5.0 8.1 8.5 2.3 6.1 4.9 8.5 7.0 1.7 5.8

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables TABLE 10-9

119

Dietary Amino Acid Requirements of Lactating Sows (90% dry matter) a Sow Postfarrowing Weight (kg) 175 175 175 Anticipated Lactational Weight Change (kg) b

DE content of diet (kcal/kg) ME content of diet (kcal/kg) c Estimated DE intake (kcal/day) Estimated ME intake (kcal/day) c Estimated feed intake (kg/day) Crude protein (%) d

175

175

175

0 0 Daily Weight Gain of Pigs (g) b 150 200

0

110

110

110

250

150

200

250

3,400 3,265 14,645 14,060 4.31 16.3

3,400 3,265 21,765 20,895 6.40 18.4

3,400 3,265 12,120 11,635 3.56 17.2

3,400 3,265 15,680 15,055 4.61 18.5

3,400 3,265 19,240 18,470 5.66 19.2

0.44 0.34 0.48 0.97 0.85 0.22 0.42 0.46 0.97 0.53 0.16 0.73

0.50 0.36 0.50 1.03 0.90 0.23 0.43 0.49 1.02 0.56 0.17 0.77

3,400 3,265 18,205 17,475 5.35 17.5

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.36 0.28 0.40 0.80 0.71 0.19 0.35 0.39 0.80 0.45 0.13 0.60

0.44 0.32 0.44 0.90 0.79 0.21 0.39 0.43 0.89 0.49 0.14 0.67

Amino acid requirements True ileal digestible basis (%) 0.49 0.35 0.34 0.30 0.47 0.44 0.96 0.87 0.85 0.77 0.22 0.20 0.41 0.39 0.46 0.42 0.95 0.88 0.52 0.50 0.15 0.15 0.72 0.66

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.34 0.27 0.37 0.77 0.66 0.18 0.33 0.36 0.75 0.40 0.11 0.55

0.41 0.30 0.41 0.86 0.73 0.20 0.36 0.40 0.83 0.43 0.12 0.61

Apparent ileal digestible basis (%) 0.46 0.33 0.32 0.29 0.44 0.41 0.92 0.83 0.79 0.72 0.21 0.19 0.38 0.36 0.43 0.39 0.89 0.82 0.46 0.44 0.13 0.13 0.66 0.61

0.41 0.32 0.44 0.92 0.79 0.21 0.39 0.43 0.90 0.47 0.14 0.67

0.47 0.34 0.47 0.98 0.84 0.22 0.40 0.46 0.96 0.49 0.14 0.71

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

0.40 0.32 0.45 0.86 0.82 0.21 0.40 0.43 0.90 0.54 0.15 0.68

0.48 0.36 0.50 0.97 0.91 0.23 0.44 0.48 1.00 0.58 0.16 0.76

Total basis (%) d 0.54 0.39 0.38 0.34 0.53 0.50 1.05 0.95 0.97 0.89 0.24 0.22 0.46 0.44 0.52 0.47 1.07 0.98 0.61 0.58 0.17 0.17 0.82 0.76

0.49 0.38 0.54 1.05 0.97 0.24 0.47 0.52 1.08 0.63 0.18 0.83

0.55 0.40 0.57 1.12 1.03 0.26 0.49 0.55 1.14 0.65 0.19 0.88

a

Daily intakes of DE and feed and the amino acid requirements are estimated by the lactation model. Assumes 10 pigs per litter and a 21-day lactation period. c Assumes that ME is 96% of DE. In corn-soybean meal diets of these crude protein levels, ME is 95–96% of DE. d Crude protein and total amino acid requirements are based on a corn–soybean meal diet. b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

120

Nutrient Requirements of Swine

TABLE 10-10

Daily Amino Acid Requirements of Lactating Sows (90% dry matter) a Sow Postfarrowing Weight (kg) 175 175 175 Anticipated Lactational Weight Change (kg) b

DE content of diet (kcal/kg) ME content of diet (kcal/kg)c Estimated DE intake (kcal/day) Estimated ME intake (kcal/day)c Estimated feed intake (kg/day) Crude protein (%)d

175

175

175

0 0 Daily Weight Gain of Pigs(g) b 150 200

0

110

110

110

250

150

200

250

3,400 3,265 14,645 14,060 4.31 16.3

3,400 3,265 21,765 20,895 6.40 18.4

3,400 3,265 12,120 11,635 3.56 17.2

3,400 3,265 15,680 15,055 4.61 18.5

3,400 3,265 19,240 18,470 5.66 19.2

3,400 3,265 18,205 17,475 5.35 17.5

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

15.6 12.2 17.2 34.4 30.7 8.0 15.3 16.8 34.6 19.5 5.5 25.8

23.4 17.0 23.6 48.0 42.5 11.0 20.6 23.3 47.9 26.4 7.6 35.8

Amino acid requirements True ileal digestible basis (g/day) 31.1 12.5 21.7 10.9 30.1 15.6 61.5 31.0 54.3 27.6 14.1 7.2 26.0 13.9 29.7 14.9 61.1 31.4 33.3 17.7 9.7 5.2 45.8 23.6

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

14.6 11.5 15.9 33.0 28.4 7.6 14.2 15.5 32.3 17.1 4.7 23.6

22.0 16.0 21.9 45.9 39.4 10.5 19.2 21.6 44.7 23.1 6.6 32.8

Apparent ileal digestible basis (g/day) 29.3 11.7 20.5 10.2 27.9 14.5 58.7 29.7 50.4 25.5 13.4 6.8 24.1 12.9 27.6 13.8 57.1 29.3 29.2 15.5 8.4 4.5 42.0 21.6

Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine

17.4 13.8 19.5 37.2 35.3 8.8 17.3 18.7 38.7 23.0 6.3 29.5

25.8 19.1 26.8 52.1 48.6 12.2 23.4 25.9 53.4 31.1 8.6 40.9

Total basis (g/day) d 34.3 14.0 24.4 12.2 34.1 17.7 67.0 33.7 61.9 31.6 15.6 7.9 29.4 15.7 33.2 16.6 68.2 35.1 39.1 20.8 11.0 5.9 52.3 26.9

a

Daily intakes of DE and feed and the amino acid requirements are estimated by the lactation model. Assumes 10 pigs per litter and a 21-day lactation period. c Assumes that ME is 96% of DE. In corn–soybean meal diets of these crude protein levels, ME is 95–96% of DE. d Crude protein and total amino acid requirements are based on a corn–soybean meal diet. b

Copyright © National Academy of Sciences. All rights reserved.

20.3 15.6 22.1 44.5 39.4 10.2 19.2 21.4 44.6 24.6 7.3 33.6

28.0 20.3 28.5 58.1 51.2 13.2 24.5 27.9 57.8 31.5 9.4 43.6

19.1 14.7 20.5 42.6 36.5 9.7 17.8 19.9 41.7 21.6 6.3 30.8

26.4 19.2 26.5 55.4 47.5 12.6 22.8 25.9 54.1 27.7 8.1 40.0

22.4 17.5 25.0 48.6 44.9 11.3 21.7 23.9 49.8 28.8 8.2 38.4

30.8 22.8 32.3 63.5 58.2 14.6 27.8 31.1 64.6 36.9 10.6 49.8

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables TABLE 10-11 matter) a

121

Dietary Mineral, Vitamin, and Fatty Acid Requirements of Gestating and Lactating Sows (90% dry

DE content of diet (kcal/kg) ME content of diet (kcal/kg) b DE intake (kcal/day) ME intake (kcal/day)b Feed intake (kg/day)

Gestation

Lactation

3,400 3,265 6,290 6,040 1.85

3,400 3,265 17,850 17,135 5.25 Requirements (% or amount/kg of diet)

Mineral elements Calcium (%) Phosphorus, total (%) Phosphorus, available (%) Sodium (%) Chlorine (%) Magnesium (%) Potassium (%) Copper (mg) Iodine (mg) Iron (mg) Manganese (mg) Selenium (mg) Zinc (mg) Vitamins Vitamin A (IU) c Vitamin D3 (IU) c Vitamin E (IU) c Vitamin K (menadione) (mg) Biotin (mg) Choline (g) Folacin (mg) Niacin, available (mg) d Pantothenic acid (mg) Riboflavin (mg) Thiamin (mg) Vitamin B6 (mg) Vitamin B12 (mg) Linoleic acid (%)

0.75 0.60 0.35 0.15 0.12 0.04 0.20 5.00 0.14 80 20 0.15 50

0.75 0.60 0.35 0.20 0.16 0.04 0.20 5.00 0.14 80 20 0.15 50

4,000 200 44 0.50 0.20 1.25 1.30 10 12 3.75 1.00 1.00 15

2,000 200 44 0.50 0.20 1.00 1.30 10 12 3.75 1.00 1.00 15

0.10

0.10

a

The requirements are based on the daily consumption of 1.85 and 5.25 kg of feed, respectively. If lower amounts of feed are consumed, the dietary percentage may need to be increased. b Assumes that ME is 96% of DE. c Conversions: 1 IU vitamin A 4 0.344 mg retinyl acetate; 1 IU vitamin D3 4 0.025 mg cholecalciferol; 1 IU vitamin E 4 0.67 mg of D-a-tocopherol or 1 mg of DL-atocopheryl acetate. d The niacin in corn, grain sorghum, wheat, and barley is unavailable. Similarly, the niacin in by-products made from these cereal grains is poorly available unless the byproducts have undergone a fermentation or wet-milling process.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

122

Nutrient Requirements of Swine

TABLE 10-12 matter) a

Daily Mineral, Vitamin, and Fatty Acid Requirements of Gestating and Lactating Sows (90% dry

DE content of diet (kcal/kg) ME content of diet (kcal/kg) b DE intake (kcal/day) ME intake (kcal/day) b Feed intake (kg/day)

Gestation

Lactation

3,400 3,265 6,290 6,040 1.85

3,400 3,265 17,850 17,135 5.25 Requirements (amount/day)

Mineral elements Calcium (g) Phosphorus, total (g) Phosphorus, available (g) Sodium (g) Chlorine (g) Magnesium (g) Potassium (g) Copper (mg) Iodine (mg) Iron (mg) Manganese (mg) Selenium (mg) Zinc (mg) Vitamins Vitamin A (IU) c Vitamin D3 (IU) c Vitamin E (IU) c Vitamin K (menadione) (mg) Biotin (mg) Choline (g) Folacin (mg) Niacin, available (mg) d Pantothenic acid (mg) Riboflavin (mg) Thiamin (mg) Vitamin B6 (mg) Vitamin B12 (mg) Linoleic acid (g)

13.9 11.1 6.5 2.8 2.2 0.7 3.7 9.3 0.3 148 37 0.3 93

39.4 31.5 18.4 10.5 8.4 2.1 10.5 26.3 0.7 420 105 0.8 263

7,400 370 81 0.9 0.4 2.3 2.4 19 22 6.9 1.9 1.9 28

10,500 1,050 231 2.6 1.1 5.3 6.8 53 63 19.7 5.3 5.3 79

1.9

5.3

a

The daily amounts of minerals and vitamins are based on the daily consumption of 1.85 and 5.25 kg of feed, respectively. If lower amounts of feed are consumed, the dietary percentages may need to be increased. b Assumes that ME is 96% of DE. c Conversions: 1 IU vitamin A 4 0.344 mg retinyl acetate; 1 IU vitamin D3 4 0.025 mg cholecalciferol; 1 IU vitamin E 4 0.67 mg of D-a-tocopherol or 1 mg of DL-atocopheryl acetate. d The niacin in corn, grain sorghum, wheat, and barley is unavailable. Similarly, the niacin in by-products made from these cereal grains is poorly available unless the byproducts have undergone a fermentation or wet-milling process.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Nutrient Requirement Tables

123

TABLE 10-13 Dietary and Daily Amino Acid, Mineral, Vitamin, and Fatty Acid Requirements of Sexually Active Boars (90% dry matter) a DE content of diet (kcal/kg) ME content of diet (kcal/kg) DE intake (kcal/day) ME intake (kcal/day) Feed intake (kg/day) Crude protein (%) b

3,400 3,265 6,800 6,530 2.00 13.0

3,400 3,265 6,800 6,530 2.00 13.0 Requirements

Amino acids (total basis) b Arginine Histidine Isoleucine Leucine Lysine Methionine Methionine ` cystine Phenylalanine Phenylalanine ` tyrosine Threonine Tryptophan Valine Mineral elements Calcium Phosphorus, total Phosphorus, available Sodium Chlorine Magnesium Potassium Copper Iodine Iron Manganese Selenium Zinc Vitamins Vitamin A c Vitamin D3c Vitamin E c Vitamin K (menadione) Biotin Choline Folacin Niacin, available d Pantothenic acid Riboflavin Thiamin Vitamin B6 Vitamin B12 Linoleic acid

% or amount/kg of diet

amount/day

— 0.19 0.35 0.51 0.60 0.16 0.42 0.33 0.57 0.50 0.12 0.40

% % % % % % % % % % %

— 3.8 7.0 10.2 12.0 3.2 8.4 6.6 11.4 10.0 2.4 8.0

0.75 0.60 0.35 0.15 0.12 0.04 0.20 5 0.14 80 20 0.15 50

% % % % % % % mg mg mg mg mg mg

15.0 12.0 7.0 3.0 2.4 0.8 4.0 10 0.28 160 40 0.3 100

4,000 200 44 0.50 0.20 1.25 1.30 10 12 3.75 1.0 1.0 15

IU IU IU mg mg g mg mg mg mg mg mg mg

0.1

8,000 400 88 1.0 0.4 2.5 2.6 20 24 7.5 2.0 2.0 30

%

a

2.0

g g g g g g g g g g g g g g g g g g mg mg mg mg mg mg IU IU IU mg mg g mg mg mg mg mg mg mg g

The requirements are based on the daily consumption of 2.0 kg of feed. Feed intake may need to be adjusted, depending on the weight of the boar and the amount of weight gain desired. b Assumes a corn–soybean meal diet. The lysine requirement was set as 0.60% (12.0 g/day). Other amino acids were calculated using ratios (total basis) similar to those for gestating sows. c Conversions: 1 IU vitamin A 4 0.344 mg retinyl acetate; 1 IU vitamin D3 4 0.025 mg cholecalciferol; 1 IU vitamin E 4 0.67 mg of D-a-tocopherol or 1 mg of DL-atocopheryl acetate. d The niacin in corn, grain sorghum, wheat, and barley is unavailable. Similarly, the niacin in by-products made from these cereal grains is poorly available unless the byproducts have undergone a fermentation or wet-milling process.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

11 When diets are formulated to meet the recommended nutrient requirements of swine, it is necessary to know the nutrient composition of and, if possible, the bioavailability of nutrients in each ingredient used. Individual feed ingredients may vary widely in composition because of variation in cultivars, growing conditions, processing and storage conditions, and nutrient status. Variations in analytical procedures also affect the results that are obtained. Furthermore, the amount of dry matter in the ingredients also affects their nutrient concentrations. The nutrient values given in the following tables are averages, reflecting the concentration of nutrients most likely to be present in feeds that are commonly used in swine diets. They are intended to be used only as a guide and users are encouraged to have a chemical analysis of feed ingredients performed prior to widespread use. A total of 79 feed ingredients have been included in this edition. Additional information included in the composition tables include estimates of net energy, neutral detergent fiber, acid detergent fiber, phosphorus bioavailability, and beta-carotene as well as estimates of apparent and true ileal amino acid digestibility. In many instances, values in this edition are different from those previously published. These changes reflect results of analyses of feed ingredients obtained from contemporary crop cultivars, newer processing techniques, and improved analytical procedures. The subcommittee obtained much of the data included in Tables 11-1 and 11-2 from United States–Canadian Tables of Feed Composition (National Research Council, 1982), International Feeds Institute Tables of Feed Composition (Fonnesbeck et al., 1984), Centraal Veevoederbureau’s Veevoedertabel, (Centraal Veevoederbureau, 1994), L’Alimentation des Animaux Monogastriques (Institut National de la Recherche Agronomique, 1984), Feedstuffs Ingredient Analysis Table (Dale, 1995), Nutrient Requirements of Poultry (National Research Council, 1993), UK Tables of Nutritive Value and Chemical Composition of

Composition of Feed Ingredients

Feedstuffs (Ministry of Agriculture Fisheries and Food Standing Committee on Tables of Feed Composition, 1990), Raw Material Compendium (Novus, 1994), and ˆ Rhodimet Nutrition Guide (Rhone-Poulenc, 1993b). Additional information was provided by suppliers of relatively new ingredients (e.g., Growmark Inc., for potato protein concentrate; American Protein Corporation, Merricks, and DuCoa for spray dried animal plasma and animal blood cells; Milk Specialities Company for whey permeate; International Ingredient Corporation for milk-based products) as well as individual scientists. Energy values were obtained from a summary compiled by R. C. Ewan of Iowa State University (Ewan, 1996) and from Noblet and Henry (1991) and Noblet et al. (1993; 1994). Phosphorus bioavailability estimates are largely based on data from Cromwell (1992) and Jongbloed (1987). Vitamin levels for Table 11-3 were largely obtained from the same sources listed for Table 11-1 and 11-2, with the exception of biotin, folic acid, b-carotene, and vitamin E, which were obtained from various publications from Hoffman–La Roche (Roche, 1986; 1987a,b; 1992). Two publications by Frigg (Frigg, 1984; Frigg and Volker, 1994) provided additional values for biotin content. Vitamin E values of feedstuffs are dramatically lower than previous publications because only (a–tocopherol values obtained by highperformance liquid chromatography were included (Cort et al., 1983). The values for the amino acid composition of feedstuffs in Table 11-4 were largely obtained from Degussa’s book The Amino Acid Composition of Feedstuffs (Fickler et al., 1995), ADM BioProducts Amino Acid Database (Archer Daniels Midland Company, 1995), Heartland Lysine’s Apparent Ileal Digestibility of Crude Protein and Essential Amino Acids in Feedstuffs for Swine—1995 (Heartland Lysine, 1995), ˆ BioKyowa’s amino acid data base (Southern, 1991), Rho ne-Poulenc’s Rhodimet Nutrition Guide ˆ (Rhone-Poulenc, 1993a), Eurolysine’s Ileal Digestibility of

124

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients Amino Acids in Feedstuffs for Pigs (Jondreville et al., 1995), comprehensive studies of the North Central Region Committee on Swine Nutrition (North Central Region Committee on Swine Nutrition [NCR-42], 1992; 1993; 1995), and the National Research Council’s Nutrient Requirements of Poultry (National Research Council, 1993). The Fats and Proteins Research Foundation supplied the subcommittee with a survey of the protein, amino acid, calcium, and phosphorus content of animal-based protein sources (Knabe, 1995). Knowledge of the availability of amino acids in feed ingredients is important for consistent formulation of diets that meet the pig’s amino acid requirements. The amounts of amino acids that are available to the animal are often much lower than the quantity contained in feed. Also, large variation exists among feedstuffs in their digestibilities of various amino acids. As a consequence, it is generally considered to be more accurate to formulate diets on an available or digestible basis rather than total amino acid basis. Tables 11-5 and 11-6 provide estimates of apparent and true ileal amino acid digestibilities. The values presented are mostly obtained from Heartland Lysine’s Apparent Ileal Digestibility of Crude Protein and Essential Amino Acids in Feedstuffs for Swine—1995 (Heartland Lysine, ˆ 1995), Rho ne-Poulenc’s Rhodimet Nutrition Guide ˆ (Rhone-Poulenc, 1993a), BioKyowa’s Bulletin Digestible Amino Acids and Digestible Amino Acid Requirements for Swine (Southern, 1991), and Eurolysine’s Ileal Digestibility of Amino Acids in Feedstuffs for Pigs (Jondreville et al., 1995). Many factors influence the amino acid composition of grains and protein supplements. For accurate and economical feed formulation, it is desirable to know the amino acid composition of the actual ingredient to be used in the

125

diet. However, it is generally not feasible to analyse all samples of feed ingredients prior to their use. Research has been conducted at several laboratories using regression analysis to estimate the amino acid composition of the feed from its proximate composition. Equations for estimating the amino acid content of feedstuffs based on the protein content are presented in Table 11-7 and were obtained from Degussa’s book The Amino Acid Composition of Feedstuffs (Fickler et al., 1995). However, caution should be taken in using this procedure because for some amino acids (e.g., lysine in corn), the correlation between amino acids and crude protein is low. Mineral concentrations of macro mineral sources shown in Table 11-8 were obtained primarily from International Feeds Institute Tables of Feed Composition (Fonnesbeck et al., 1984), Nutrient Requirements of Poultry (National Research Council, 1993), and Macrominerals (Axe, 1994), as well as from suppliers of mineral supplements including Consolidated Minerals Inc. (Plant City, FL), J. R. Simplot Company (Lathrop, CA), Mallinckrodt (IMC-Agrico, Bannockburn, IL), Nutra-Flo Company (Sioux City, IA), White Springs Agricultural Chemicals (White Springs, FL), Occidental Chemical Corporation (White Springs, FL) and PCS Phosphate Company (Raleigh, NC). Trace mineral concentrations in Table 11-9 are largely from National Research Council (1982), Fonnesbeck et al. (1984), Ammerman et al. (1995), Reese et al. (1995), Bernhardt (1996), and Nelson (1995). The fatty acid composition of various sources of fats and oils presented in Table 11-10 are from Feeding Fats published by the Fats and Proteins Research Foundation (Pearl, 1995) and USDA Food Composition Standard Release 11-1 (U.S. Department of Agriculture, 1997). Energy values for fats are from the summary of Ewan (1996) and the work of Powles et al. (1995).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

126

Nutrient Requirements of Swine

TABLE 11-1 Entry Number 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Chemical Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis) a

Description Alfalfa meal dehydrated, 17% CP meal dehydrated, 20% CP Bakery Waste dried bakery product Barley grain, two row grain, six row grain, hulless Beet, Sugar pulp, dried Blood meal, conventional meal, flash dried meal, spray or ring dried plasma, spray driede cells, spray driede Brewers’ Grain dried Buckwheat, Common grain Canola (Rapeseed) meal, sol. extr. Casein dried Cassava (Tapioca or Manioc) meal, dehydrated Coconut (Copra) meal, sol. extr. Corn, Yellow distillers’ grain distillers’ grain with solubles distillers’ solubles gluten feed gluten meal, 60% CP grain grits by-product (Hominy Feed) Cottonseed meal, mech. extr., 41% CP meal, sol. extr. 41% CP Fababean (Broadbean) seeds Feather meal, hydrolyzed Fish Anchovy meal, mech. extr. Herring meal, mech. extr. Menhaden meal, mech. extr. White meal, mech. extr. solubles, condensed solubles, dried Flax (Linseed) meal, sol. extr. Lentil seeds Lupin (Sweet White) seeds Meat meal rendered meal rendered with bone Milk (Cattle) skim, dried Millet (Proso) grain

International Feed Numberb

Dry Matter (%)

DE (kcal/kg)

ME (kcal/kg)

NE (kcal/kg)

Crude Protein (%)

Crude Fat (%)

Linoleic Acid (%)

ADF (%)

Calcium (%)

Phosphorus (%)

Bioavailability of Phosphorus c (%)

NDF (%)

1-00-023 1-00-024

92 92

1,830 2,095

1,650 1,885

910 1,290

17.0 19.6

2.6 3.3

0.35 0.44

41.2 38.8

30.2 26.4

1.53 1.61

0.26 0.28

100 —

4-00-466

91

3,940

3,700

2,415 d

10.8

11.3

5.70

2.0

1.3

0.13

0.25



4-00-572 4-00-574 4-00-552

89 89 88

3,050 3,050 3,360

2,910 2,910 3,320

2,340 2,310 2,650

11.3 10.5 14.9

1.9 1.9 2.1

0.88 0.91 1.14

18.0 18.6 10.1

6.2 7.0 2.2

0.06 0.06 0.04

0.35 0.36 0.45

— 30 —

4-00-669

91

5-00-380 5-26-006 5-00-381 — —

92 92 93 91 92

2,865

2,495

1,860

8.6

2,850 2,300 3,370 — —

2,350 1,950 2,945 — —

1,950 1,385 d 2,070 — —

77.1 87.6 88.8 78.0 92.0

0.8



42.4

24.3

0.70

0.10



1.6 1.6 1.3 2.0 1.5

0.09 — 0.17 — —

13.6 — — — —

1.8 — — — —

0.37 0.21 0.41 0.15 0.02

0.27 0.21 0.30 1.71 0.37

— — 92 — —

5-02-141

92

2,100

1,960

1,630

26.5

7.3

3.14

48.7

21.9

0.32

0.56

34

4-00-994

88

2,825

2,640

1,620

11.1

2.4

0.53

17.8

14.3

0.09

0.31

5-06-145

90

2,885

2,640

1,610

35.6

3.5

0.42

21.2

17.2

0.63

1.01

5-01-162

91

4,135

3,535

2,555

88.7

0.8

0.03





0.61

0.82



4-01-152

88

3,385

3,330

2,330

3.3

0.5



5-01-573

92

3,010

2,565

1,695

21.9

3.0

5-02-842 5-02-843 5-02-844 5-02-903 5-28-242 4-02-935 4-03-011

94 93 92 90 90 89 90

3,100 3,200 3,325 2,990 4,225 3,525 3,355

2,715 2,820 2,945 2,605 3,830 3,420 3,210

1,170 d 2,065 2,250 1,740 2,550 2,395 2,260

24.8 27.7 26.7 21.5 60.2 8.3 10.3

5-01-617 5-07-872

92 90

2,945 2,575

2,690 2,315

1,870 1,325

5-09-262

87

3,245

3,045

5-03-795

93

2,990

5-01-985 5-02-000 5-02-009 5-02-025 5-01-969 5-01-971

92 93 92 91 51 92

5-02-048

— 21

7.7

4.6

0.22

0.13



0.03

51.3

25.5

0.16

0.58



7.9 8.4 9.1 3.0 2.9 3.9 6.7

4.46 2.15 5.36 1.43 1.17 1.92 2.97

40.4 34.6 24.8 33.3 8.7 9.6 28.5

17.5 16.3 7.5 10.7 4.6 2.8 8.1

0.10 0.20 0.29 0.22 0.05 0.03 0.05

0.40 0.77 1.03 0.83 0.44 0.28 0.43

— 77 — 59 15 14 14

42.4 41.4

6.1 1.5

3.15 0.51

25.7 28.4

18.0 19.4

0.23 0.19

1.03 1.06

— 1

2,000

25.4

1.4

0.62

13.7

9.7

0.11

0.48



2,485

2,250

84.5

4.6

0.83





0.33

0.50

31

3,230 3,960 3,770 3,395 1,910 3,310

2,695 3,260 3,360 2,810 1,625 3,045

1,695 d 2,020 2,335 2,020 995 d 1,770

64.6 68.1 62.3 63.3 32.7 64.2

7.9 9.2 9.4 4.8 5.6 7.4

0.27 0.15 0.12 0.08 — 0.12

— — — — — —

— — — — — —

3.93 2.40 5.21 6.65 0.22 0.55

2.55 1.76 3.04 3.59 0.59 1.25

— — 94 — — —

90

3,060

2,710

1,840

33.6

1.8

0.36

23.9

15.0

0.39

0.83



5-02-506

89

3,540

3,450

2,205

24.4

1.3

0.41

10.1

5.4

0.10

0.38



5-27-717

89

3,450

3,305

2,130

34.9

9.2

1.62

20.3

16.7

0.22

0.51



5-00-385 5-00-388

94 93

2,695 2,440

2,595 2,225

2,175 1,355

54.0 51.5

12.0 10.9

0.80 0.72

31.6 32.5

8.3 5.6

7.69 9.99

3.88 4.98

— f 90

5-01-175

96

3,980

3,715

2,360

34.6

0.9

0.01





1.31

1.00

91

4-03-120

90

3,020

2,950

2,095

11.1

3.5

1.92

15.8

13.8

0.03

0.31

— Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-1 Entry Number

(continued)

Description Oat grain grain, naked groat Pea seeds Peanut (Groundnut) meal, mech. extr. meal, sol. extr. Potato protein concentrate Poultry by-product, meal rendered Rice bran grain, polished and broken (Brewers’ Rice) polishings Rye grain Safflower meal, sol. extr. meal without hulls, sol. extr. Sesame meal, mech. extr. Sorghum grain Soybean meal, sol. extr. meal without hulls, sol. extr. protein concentrate protein isolate seeds, heat processed Sunflower meal, sol. extr. meal without hulls, sol. extr. Triticale grain Wheat bran grain, hard red spring grain, hard red winter grain, soft red winter grain, soft white winter middlings, , 9.5% fiber red dog, , 4% fiber shorts, , 7% fiber Whey dried low lactose, dried permeate, dried Yeast, Brewers’ dried Yeast, Torula dried

43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79

127

International Feed Numberb

Dry Matter (%)

DE (kcal/kg)

ME (kcal/kg)

NE (kcal/kg)

Crude Protein (%)

Crude Fat (%)

Linoleic Acid (%)

4-03-309 4-25-101 4-03-331

89 86 90

2,770 3,480 3,690

2,710 3,410 3,465

1,760 2,160 d 2,310

5-03-600

89

3,435

3,210

5-03-649 5-03-650

92 92

3,895 3,415

5-25-392

91

5-03-798

93

4-03-928

ADF (%)

Calcium (%)

Phosphorus (%)

Bioavailability of Phosphorus c (%)

NDF (%)

11.5 17.1 13.9

4.7 6.5 6.2

1.62 2.52 2.40

27.0 9.9 —

13.5 3.7 —

0.07 0.08 0.08

0.31 0.38 0.41

22 — 13

2,195

22.8

1.2

0.47

12.7

7.2

0.11

0.39



3,560 3,245

2,280 2,170

43.2 49.1

6.5 1.2

1.73 0.30

14.6 16.2

9.1 12.2

0.17 0.22

0.59 0.65

— 12

4,140

3,880

2,040

73.8

1.7



0.17

0.19



3,090

2,860

1,945 d

64.1

12.6

2.54





4.46

2.41



90

3,100

2,850

2,040

13.3

13.0

4.12

23.7

13.9

0.07

1.61

4-03-932 4-03-943

89 90

3,565 3,770

3,350 3,350

2,295 2,070d

7.9 13.0

1.0 13.7

0.28 3.58

12.2 —

3.1 4.0

0.04 0.09

0.18 1.18

— —

4-04-047

88

3,270

3,060

2,300

11.8

1.6

0.76

12.3

4.6

0.06

0.33



5-04-110 5-07-959

92 92

2,840 3,055

2,170 2,910

870 1,585

23.4 42.5

1.4 1.3

0.84 0.74

55.9 25.9

38.8 18.0

0.34 0.37

0.75 1.31

— —

5-04-220

93

3,350

3,035

2,090

42.6

7.5

3.07

18.0

13.2

1.90

1.22



4-20-893

89

3,380

3,340

2,255

9.2

2.9

1.13

18.0

8.3

0.03

0.29

20

5-04-604 5-04-612 — 5-08-038 5-04-597

89 90 90 92 90

3,490 3,685 4,100 4,150 4,140

3,180 3,380 3,500 3,560 3,690

1,935 2,020 2,000 d 2,000 2,880

43.8 47.5 64.0 85.8 35.2

1.5 3.0 3.0 0.6 18.0

0.69 0.60 — — 9.13

13.3 8.9 — — 13.9

9.4 5.4 — — 8.0

0.32 0.34 0.35 0.15 0.25

0.65 0.69 0.81 0.65 0.59

31 23 — — —

5-09-340 5-04-739

90 93

2,010 2,840

1,830 2,735

1,230 1,635

26.8 42.2

1.3 2.9

0.98 1.07

42.4 27.8

30.3 18.4

0.36 0.37

0.86 1.01

3 —

4-20-362

90

3,320

3,180

2,420

12.5

1.8

0.71

12.7

3.8

0.05

0.33

46

4-05-190 4-05-258 4-05-268 4-05-294 4-05-337 4-05-205 4-05-203 4-05-201

89 88 88 88 89 89 88 88

2,420 3,400 3,365 3,450 3,400 3,075 3,140 2,985

2,275 3,250 3,210 3,305 3,285 3,025 2,925 2,820

1,400 1,925 d 2,225 2,400 2,375 1,560 2,090 2,120

15.7 14.1 13.5 11.5 11.8 15.9 15.3 16.0

4.0 2.0 2.0 1.9 2.1 4.2 3.3 4.6

1.80 — 0.93 — 0.83 1.74 — 1.90

42.1 — 13.5 — 12.0 35.6 18.7 28.4

13.0 — 4.0 — 3.7 10.7 4.3 8.6

0.16 0.05 0.06 0.04 0.05 0.12 0.07 0.09

1.20 0.36 0.37 0.39 0.35 0.93 0.57 0.84

29 — 50 50 — 41 — —

4-01-182 4-01-186 —

96 96 96

3,335 3,045 3,435

3,190 2,910 3,300

2,215 2,030 2,260 d

12.1 17.6 3.8

0.9 1.1 0.2

0.01 0.04 —

— — —

— — —

0.75 2.00 0.86

0.72 1.37 0.66

97 — —

7-05-527

93

3,325

3,025

2,075

45.9

1.7

0.04

0.16

1.44



7-05-534

93

3,110

2,765

1,985

46.4

2.4

0.05

0.58

1.52





a

1.8

4.0 —

3.0 —

25

Dash indicates that no data were available. First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other five digits are the International Feed Number. c Estimated bioavailability, relative to that in monosodium or monocalcium phosphate. d Based on chemical composition using Equation 1-12 in Chapter 1. e DE, ME, and NE of spray-dried plasma and spray-dried blood cells have not been determined experimentally and composition data are insufficient to accurately determine calculated values. f Some meat and bone meals may have phosphorus bioavailabilities of 70% or less. b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

128

Nutrient Requirements of Swine

TABLE 11-2

Mineral Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis)a

Entry Number Description 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Alfalfa meal dehydrated, 17% CP meal dehydrated, 20% CP Bakery Waste dried bakery product Barley grain, two row grain, six row grain, hulless Beet, Sugar pulp, dried Blood meal, conventional meal, flash dried meal, spray or ring dried plasma, spray dried cells, spray dried Brewers’ Grains dried Buckwheat, Common grain Canola (Rapeseed) meal, sol. extr. Casein dried Cassava (Tapioca or Manioc) meal, dehydrated Coconut (Copra) meal, sol. extr. Corn, Yellow distillers’ grain distillers’ grains with solubles distillers’ solubles gluten feed gluten meal, 60% CP grain grits by-product (Hominy Feed) Cottonseed meal, mech. extr. 41% CP meal, sol. extr. 41% CP Fababean (Broadbean) seeds Feather meal, hydrolyzed Fish Anchovy meal, mech. extr. Herring meal, mech. extr. Menhaden meal, mech. extr. White meal, mech. extr. solubles, condensed solubles, dried Flax (Linseed) meal, sol. extr. Lentil seeds Lupin (Sweet White) seeds Meat meal rendered meal rendered with bone Milk (Cattle) skim, dried Millet (Proso) grain

International Dry Feed Matter Number b (%)

Calcium (%)

Phosphorus (%)

Sodium (%)

Chlorine (%)

Potassium (%)

Magnesium (%)

Sulfur (%)

1-00-023 1-00-024

92 92

1.53 1.61

0.26 0.28

0.09 0.09

0.47 0.47

2.30 2.40

0.23 0.36

0.29 10 0.26 11

4-00-466

91

0.13

0.25

1.14

1.48

0.39

0.24

0.02

4-00-572 4-00-574 4-00-552

89 89 88

0.06 0.06 0.04

0.35 0.36 0.45

0.04 0.02 0.02

0.12 0.15 0.10

0.45 0.47 0.44

0.14 0.12 0.12

0.15 0.15 0.12

4-00-669

91

0.70

0.10

0.20

0.10

0.61

0.22

5-00-380 5-26-006 5-00-381 — —

92 92 93 91 92

0.37 0.21 0.41 0.15 0.02

0.27 0.21 0.30 1.71 0.37

0.50 0.29 0.44 3.02 0.58

0.30 0.38 0.25 1.50 1.40

0.11 0.14 0.15 0.20 0.62

0.11 0.21 0.11 0.34 —

5-02-141

92

0.32

0.56

0.26

0.15

0.08

4-00-994

88

0.09

0.31

0.05

0.05

5-06-145

90

0.63

1.01

0.07

5-01-162

91

0.61

0.82

4-01-152

88

0.22

5-01-573

92

5-02-842 5-02-843 5-02-844 5-02-903 5-28-242 4-02-935 4-03-011

Copper (mg/ kg)

Iron (mg/ kg)

Manganese (mg/ kg)

Selenium c (mg/ kg)

Zinc (mg/ kg)

333 346

32 42

0.34 0.29

24 21

5

28

65



15

7 8 5

78 88 56

18 16 16

0.19 0.10 —

25 15 27

0.31 11

411

46

0.09

12

0.48 11 0.45 6 0.47 8 — — — —

1,922 2,341 2,919 55 2,700

6 10 6

0.58 — — — —

38 16 30 — —

0.16

0.31 21

250

38

0.70

62

0.41

0.09

0.14 10

44

34

0.18

9

0.11

1.22

0.51

0.85

6

142

49

1.10

69

0.01

0.04

0.01

0.01

0.60

4

14

4

0.16

30

0.13

0.03

0.07

0.49

0.11

0.50

4

18

28

0.10

10

0.16

0.58

0.04

0.37

1.83

0.31

0.31 25

486

69



49

94 93 92 90 90 89 90

0.10 0.20 0.29 0.22 0.05 0.03 0.05

0.40 0.77 1.03 0.83 0.44 0.28 0.43

0.09 0.25 0.26 0.15 0.02 0.02 0.08

0.08 0.20 0.25 0.22 0.06 0.05 0.07

0.17 0.84 1.50 0.98 0.18 0.33 0.61

0.25 0.19 0.64 0.33 0.08 0.12 0.24

0.43 0.30 0.37 0.22 0.43 0.13 0.03

45 57 83 48 26 3 13

220 257 560 460 282 29 67

22 24 74 24 4 7 15

0.40 0.39 0.33 0.27 1.00 0.07 0.10

55 80 85 70 33 18 30

5-01-617 5-07-872

92 90

0.23 0.19

1.03 1.06

0.04 0.04

0.04 0.05

1.34 1.40

0.52 0.50

0.40 19 0.31 18

160 184

23 20

0.90 0.80

64 70

5-09-262

87

0.11

0.48

0.03

0.07

1.20

0.15

0.29 11

75

15

0.02

42

5-03-795

93

0.33

0.50

0.34

0.26

0.19

0.20

1.39 10

76

10

0.69

111

5-01-985 5-02-000 5-02-009 5-02-025 5-01-969 5-01-971

92 93 92 91 51 92

3.93 2.40 5.21 6.65 0.22 0.55

2.55 1.76 3.04 3.59 0.59 1.25

0.88 0.61 0.40 0.78 0.21 0.37

1.02 1.12 0.55 1.28 2.70 6.29

0.75 1.01 0.70 0.85 1.61 2.03

0.24 0.18 0.16 0.18 0.02 0.30

0.77 9 0.69 6 0.45 11 0.48 6 0.12 45 0.40 35

220 181 440 299 160 300

10 8 37 12 14 50

1.36 1.93 2.10 1.62 2.00 2.20

103 132 147 90 38 76

5-02-048

90

0.39

0.83

0.13

0.06

1.26

0.54

0.39 22

270

41

0.63

66

5-02-506

89

0.10

0.38

0.02

0.03

0.89

0.12

0.20 10

85

13

0.10

25

5-27-717

89

0.22

0.51

0.02

0.03

1.10

0.19

0.24

6

54

1,390

0.07

32

5-00-385 5-00-388

94 93

9 9.99

3.88 4.98

0.80 0.63

0.97 0.69

0.57 0.65

0.35 0.41

0.45 10 0.38 11

440 606

10 17

0.37 0.31

94 96

5-01-175

96

1.31

1.00

0.48

1.00

1.60

0.12

0.32

5

8

2

0.12

42

4-03-120

90

0.03

0.31

0.04

0.03

0.43

0.16

0.14 26

71

30

0.70

— —

18 Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-2

(continued)

Entry Number Description 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79

129

Oat grain grain, naked groat Pea seeds Peanut (Groundnut) meal, mech. extr. meal, sol. extr. Potato protein concentrate Poultry by-product, meal rendered Rice bran grain, polished and broken (Brewers’ Rice) polishings Rye grain Safflower meal, sol. extr. meal without hulls, sol. extr. Sesame meal, mech. extr. Sorghum grain Soybean meal, sol. extr. meal without hulls, sol. extr. protein concentrate protein isolate seeds, heat processed Sunflower meal, sol. extr. meal without hulls, sol. extr. Triticale grain Wheat bran grain, hard red spring grain, hard red winter grain, soft red winter grain, soft white winter middlings, , 9.5% fiber red dog, , 4% fiber shorts, , 7% fiber Whey dried low lactose, dried permeate, dried Yeast, Brewers’ dried Yeast, Torula dried

International Dry Feed Matter Number b (%)

Calcium (%)

Phosphorus (%)

Sodium (%)

Chlorine (%)

Potassium (%)

Magnesium (%)

Sulfur (%)

Selenium c (mg/ kg)

Zinc (mg/ kg)

4-03-309 4-25-101 4-03-331

89 86 90

0.07 0.08 0.08

0.31 0.38 0.41

0.08 0.02 0.05

0.10 0.11 0.09

0.42 0.36 0.38

0.16 0.12 0.11

0.21 0.14 0.20

6 4 6

85 58 49

43 37 32

0.30 0.09 —

38 34 —

5-03-600

89

0.11

0.39

0.04

0.05

1.02

0.12

0.20

9

65

23

0.38

23

5-03-649 5-03-650

92 92

0.17 0.22

0.59 0.65

0.06 0.07

0.03 0.04

1.20 1.25

0.33 0.31

0.29 15 0.30 15

285 260

39 40

0.28 0.21

47 41

5-25-392

91

0.17

0.19

0.03

0.20

0.80

0.05

5-03-798

93

4.46

2.41

0.49

0.49

0.53

0.18

0.23 13

40

5

1.00

25

0.52 10

442

9

0.88

94

4-03-928

90

0.07

1.61

0.03

0.07

1.56

0.90

0.18

9

190

228

0.40

30

4-03-932 4-03-943

89 90

0.04 0.09

0.18 1.18

0.04 0.06

0.07 0.11

0.13 1.11

0.11 0.65

0.06 21 0.17 6

18 160

12 12

0.27 —

17 26

4-04-047

88

0.06

0.33

0.02

0.03

0.48

0.12

0.15

7

60

58

0.38

31

5-04-110 5-07-959

92 92

0.34 0.37

0.75 1.31

0.05 0.04

0.08 0.16

0.76 1.00

0.35 1.02

0.13 10 0.20 9

495 484

18 39

— —

41 33

5-04-220

93

1.90

1.22

0.04

0.07

1.10

0.54

0.56 34

93

53

0.21

100

4-20-893

89

0.03

0.29

0.01

0.09

0.35

0.15

0.08

5

45

15

0.20

15

5-04-604 5-04-612 — 5-08-038 5-04-597

89 90 90 92 90

0.32 0.34 0.35 0.15 0.25

0.65 0.69 0.81 0.65 0.59

0.01 0.02 0.05 0.07 0.03

0.05 0.05 — 0.02 0.03

1.96 2.14 2.20 0.27 1.70

0.27 0.30 0.32 0.08 0.28

0.43 0.44 — 0.71 0.30

20 20 13 14 16

202 176 110 137 80

29 36 — 5 30

0.32 0.27 — 0.14 0.11

50 55 30 34 39

5-09-340 5-04-739

90 93

0.36 0.37

0.86 1.01

0.02 0.04

0.10 0.13

1.07 1.27

0.68 0.75

0.30 26 0.38 25

254 200

41 35

0.50 0.32

66 98

4-20-362

90

0.05

0.33

0.03

0.03

0.46

0.10

0.15

8

31

43



32

4-05-190 4-05-258 4-05-268 4-05-294 4-05-337 4-05-205 4-05-203 4-05-201

89 88 88 88 89 89 88 88

0.16 0.05 0.06 0.04 0.05 0.12 0.07 0.09

1.20 0.36 0.37 0.39 0.35 0.93 0.57 0.84

0.04 0.02 0.01 0.01 0.01 0.05 0.04 0.02

0.07 0.09 0.06 0.08 0.07 0.04 0.10 0.04

1.26 0.41 0.49 0.46 0.44 1.06 0.63 1.06

0.52 0.16 0.13 0.11 0.15 0.41 0.16 0.25

0.22 14 0.17 7 0.15 6 0.16 8 0.18 7 0.17 10 0.24 6 0.20 12

170 64 39 32 60 84 46 100

113 42 34 38 37 100 55 89

0.51 0.30 0.33 0.28 0.26 0.72 0.30 0.75

100 43 40 47 28 92 65 100

4-01-182 4-01-186 —

96 96 96

0.75 2.00 0.86

0.72 1.37 0.66

0.94 1.85 1.00

1.40 3.43 2.23

1.96 4.68 2.10

0.13 0.25 —

0.72 13 1.59 3 — —

130 85 —

3 8 —

0.12 0.06 —

10 11 —

7-05-527

93

0.16

1.44

0.10

0.12

1.80

0.23

0.40 33

215

8

1.00

49

7-05-534

93

0.58

1.52

0.07

0.12

1.94

0.20

0.55 17

222

13

0.02

99

a

Copper (mg/ kg)

Iron (mg/ kg)

Manganese (mg/ kg)

Dash indicates that no data were available. First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other five digits are the International Feed Number. c Selenium values are extremely dependent on soil conditions and some values may differ substantially from those presented here. b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

130

Nutrient Requirements of Swine

TABLE 11-3 Entry Number 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Vitamin Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis) a

Description Alfalfa meal dehydrated, 17% CP meal dehydrated, 20% CP Bakery Waste dried bakery product Barley grain, two row grain, six row grain, hulless Beet, Sugar pulp, dried Blood meal, conventional meal, flash dried meal, spray or ring dried plasma, spray dried cells, spray dried Brewers’ Grains dried Buckwheat, Common grain Canola (Rapeseed) meal, sol. extr. Casein dried Cassava (Tapioca or Manioc) meal, dehydrated Coconut (Copra) meal, sol. extr. Corn, Yellow distillers’ grains distillers’ grains with solubles distillers’ solubles gluten feed gluten meal, 60% CP grain grits by-product (Hominy Feed) Cottonseed meal, mech. extr. 41% CP meal, sol. extr. 41% CP Fababean (Broadbean) seeds Feather meal, hydrolyzed Fish Anchovy meal, mech. extr. Herring meal, mech. extr. Menhaden meal, mech. extr. White meal, mech. extr. solubles, condensed solubles, dried Flax (Linseed) meal, sol. extr. Lentil seeds Lupin (Sweet White) seeds Meat meal rendered meal rendered with bone Milk (Cattle) skim, dried Millet (Proso) grain

Thiamin (mg/ kg)

Vitamin B6 (mg/ kg)

Vitamin Ed (mg/ kg)

BetaCarotene e (mg/ kg)

13.6 15.2

3.4 5.8

6.5 8.0

0 0

49.8 49.8

94.6 94.6

8.3

1.4

2.9

4.3

0



4.2

55 48 48

8.0 7.0 6.8

1.8 1.6 1.8

4.5 4.0 4.3

5.0 2.9 5.6

0 0 0

7.4 7.4 6.0

4.1 4.1 —

18

1.3

0.7

0.4

1.9

0

13.2

10.6

0.10 0.10 0.40 — —

31 23 23 — —

2.0 1.0 3.7 — —

2.4 1.4 3.2 — —

0.4 1.0 0.3 — —

4.4 4.4 4.4 — —

44 44 — — —

1.0 1.0 1.0 — —

— — — — —

1,723

7.10

43

8.0

1.4

0.6

0.7

0



0.06

440

0.64

19

12.0

5.5

4.0

3.0

0





90

0.98

6,700

0.83

160

9.5

5.8

5.2

7.2

0

13.4



5-01-167

91

0.04

205

0.51

1

2.7

1.5

0.4

0.4





4-01-152

88

0.05





3

0.3

0.8

1.6

0.7

5-01-573

92

0.25

1,089

0.30

28

6.5

3.5

0.7

4.4

5-02-842 5-02-843 5-02-844 5-02-903 5-28-242 4-02-935 4-03-011

94 93 92 90 90 89 90

0.49 0.78 1.66 0.14 0.15 0.06 0.13

1,180 2,637 4,842 1,518 330 620 1,155

0.90 0.90 1.10 0.28 0.13 0.15 0.21

37 75 116 66 55 24c 47

11.7 14.0 21.0 17.0 3.5 6.0 8.2

5.2 8.6 17.0 2.4 2.2 1.2 2.1

1.7 2.9 6.9 2.0 0.3 3.5 8.1

4.4 8.0 8.8 13.0 6.9 5.0 11.0

0 0 3 0 0 0 0

12.9 — — 8.5 6.7 8.3 6.5

3.0 3.5 — 1.0 — 0.8 9.0

5-01-617 5-07-872

92 90

0.30 0.30

2,753 2,933

1.65 1.65

38 40

10.0 12.0

5.1 5.9

6.4 7.0

5.3 5.1

0 0

35.0 14.0

0.2 0.2

5-09-262

87

0.09

1,670



26

3.0

2.9

5.5

0

0.8



5-03-795

93

0.13

891

0.20

21

10.0

2.1

0.1

3.0

78

7.3



5-01-985 5-02-000 5-02-009 5-02-025 5-01-969 5-01-971

92 93 92 91 51 92

0.13 0.13 0.13 0.13 0.18 0.26

4,408 5,306 3,056 3,099 3,519 5,507

0.37 0.37 0.37 0.37 0.02 0.60

100 93 55 59 169 271

15.0 17.0 9.0 9.9 35.0 55.0

7.1 9.9 4.9 9.1 14.6 15.6

0.3 0.4 0.5 1.7 5.5 7.4

4.0 4.8 4.0 5.9 12.2 23.8

280 403 143 90 347 401

5.0 15.0 5.0 5.0 — —

— — — — — —

5-02-048

90

0.41

1,512

1.30

33

14.7

2.9

7.5

6.0

0

2.0

0.2

5-02-506

89

0.13



0.70

22

14.9

2.4

3.9

5.5

0

0.0

1.0

5-27-717

89

0.05





5-00-385 5-00-388

94 93

0.08 0.08

2,077 1,996

0.50 0.41

57 49

5.0 4.1

4.7 4.7

0.6 0.4

2.4 4.6

5-01-175

96

0.25

1,393

0.47

12

36.4

19.1

3.7

4-03-120

90

0.16

440

0.23

23

11.0

3.8

7.3

International Feed Number b

Dry Matter (%)

1-00-023 1-00-024

Pantothenic Acid (mg/ kg)

Riboflavin (mg/ kg)

38 45

29.0 34.0

0.20

26

0.31 0.40 0.62 —

Biotin (mg/ kg)

Choline (mg/ kg)

Folacin (mg/ kg)

92 92

0.54 0.54

1,401 1,419

4.36 4.36

4-00-466

91

0.07

923

4-00-572 4-00-574 4-00-552

89 89 88

0.14 0.15 0.07

1.034 1,034 —

4-00-669

91



5-00-380 5-26-006 5-00-381 — —

92 92 93 91 92

0.03 0.08 0.28 — —

852 781 485 — —

5-02-141

92

0.24

4-00-994

88

5-06-145

818

Niacin c (mg/ kg)













Vitamin B12 (mg/ kg)

— 0 —



0.2

0.2



7.7



7.5



80 90

1.2 1.6

— —

4.1

36

4.1



5.8

0



— Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-3 Entry Number 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79

131

(continued)

Description Oat grain grain, naked groat Pea seeds Peanut (Groundnut) meal, mech. extr. meal, sol. extr. Potato protein concentrate Poultry by-product, meal rendered Rice bran grain, polished and broken (Brewers’ Rice) polishings Rye grain Safflower meal, sol. extr. meal without hulls, sol. extr. Sesame meal, mech. extr. Sorghum grain Soybean meal, sol. extr. meal without hulls, sol. extr. protein concentrate protein isolate seeds, heat processed Sunflower meal, sol. extr. meal without hulls, sol. extr. Triticale grain Wheat bran grain, hard red spring grain, hard red winter grain, soft red winter grain, soft white winter middlings, , 9.5% fiber red dog, , 4% fiber shorts, , 7% fiber Whey dried low lactose, dried permeate, dried Yeast, Brewers’ dried Yeast, Torula dried

International Feed Number b

Dry Matter (%)

4-03-309 4-25-101 4-03-331

Pantothenic Acid (mg/ kg)

Thiamin (mg/ kg)

Vitamin B6 (mg/ kg)

1.7 1.3 1.5

6.0 5.2 6.5

2.0 9.6 1.1

18.7

1.8

4.6

166 170

47.0 53.0

5.2 7.0

7.1 5.7







47

11.1

10.5

0.2

4.4

2.20 0.20

293 25

23.0 3.3

2.5 0.4

22.5 1.4

26.0 28.0

0 0

9.7 2.0

Biotin (mg/ kg)

Choline (mg/ kg)

Folacin (mg/ kg)

89 86 90

0.24 0.12 0.20

946 1,240 1,139

0.30 0.50 0.50

19 c 20 c 14

13.0 7.1 13.4

5-03-600

89

0.15

547

0.20

31

5-03-649 5-03-650

92 92

0.35 0.39

1,848 1,854

0.70 0.50

5-25-392

91







5-03-798

93

0.09

6,029

0.50

4-03-928 4-03-932

90

0.35 0.08

1,135 1,003

Niacin c (mg/ kg)

Riboflavin (mg/ kg)



Vitamin Ed (mg/ kg)

BetaCarotene e (mg/ kg)

0 0 0

7.8 2.0 —

3.7 — —

1.0

0

0.2

1.0

7.4 6.0

0 0

2.7 2.7



Vitamin B12 (mg/ kg)

— —











— — —

4-03-943

89 90

0.37

1,237

0.20

520

47.0

1.8

19.8

27.6

0

61.0

4-04-047

88

0.08

419

0.60

19

8.0

1.6

3.6

2.6

0

9.0



5-04-110 5-07-959

92 92

1.03 1.03

820 3,248

0.50 1.60

11 22

33.9 39.1

2.3 2.4

4.6 4.5

12.0 11.3

0 0

16.0 16.0

— —

5-04-220

93

0.24

1,536



30

6.0

3.6

2.8

12.5

0

1.0

4-20-893

89

0.26

668

0.17

41 c

12.4

1.3

3.0

5.2

0

5.0

5-04-604 5-04-612 — 5-08-038 5-04-597

89 90 90 92 90

0.27 0.26 — 0.30 0.24

2,794 2,731 — 2 2,307

1.37 1.37 — 2.50 3.60

34 22 — 6 22

16.0 15.0 — 4.2 15.0

2.9 3.1 — 1.7 f 2.6

4.5 3.2 — 0.3 f 11.0

6.0 6.4 — 5.4 f 10.8

0 0 0 0

2.3 2.3 — — 18.1

5-09-340 5-04-739

90 93

1.40 1.45

3,791 3,150

1.14 1.14

264 220

29.9 24.0

3.0 3.6

3.0 3.5

11.1 13.7

0 0

9.1 9.1

— —

4-20-362

90









0.4

1.7



4-05-190 4-05-258 4-05-268 4-05-294 4-05-337 4-05-205 4-05-203 4-05-201

89 88 88 88 89 89 88 88

0.36 0.11 0.11 0.11 0.11 0.33 0.11 0.24

1,232 1,026 778 1,092 1,002 1,187 1,534 1,170

0.63 0.44 0.22 0.35 0.22 0.76 0.80 1.40

186 56 c 48 c 48 c 57 c 72 42 107

31.0 12.5 9.9 9.9 11.0 15.6 13.3 22.3

4.6 1.3 1.4 1.4 1.3 1.8 2.2 3.3

8.0 5.1 4.5 4.5 4.3 16.5 22.8 18.1

12.0 3.6 3.4 2.2 4.0 9.0 4.6 7.2

4-01-182 4-01-186 —

96 96 96

0.27 0.27 —

1,820 3,571 —

0.85 0.69 —

10 19 —

47.0 69.0 —

27.1 37.2 —

4.1 5.7 —

7-05-527

93

0.63

3,984

9.90

448

109

37.0

7-05-534

93

0.58

2,881

22.4

492

84.2

49.9

462

a







— 0 0 0 0 0 0

0.1

0.2 — 0.2 0.2 — — 1.9

0

16.5 — 11.6 — 11.6 20.1 — —

1.0 — 0.4 — 0.4 3.0 — —

4.0 4.4 —

23 25 —

0.3 0.3 —

— — —

91.8

42.8

1

10.0



6.2

36.3









Dash indicates that no data were available. b First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other five digits are the International Feed Number. c The niacin in corn, oats, sorghum, and wheat grain is totally unavailable. The bioavailability of niacin in most by-products produced from these grains is probably also low. d As a-tocopherol. e Conversion of beta-carotene to vitamin A: 1 mg of all-trans beta-carotene 4 267 IU of vitamin A or 80 mg of vitamin A alcohol (retinol) or 92 mg of retinyl acetate. f The riboflavin, thiamin, and vitamin B6 in soybean protein isolate are totally unavailable.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

132

Nutrient Requirements of Swine

TABLE 11-4

Amino Acid Composition of Some Feed Ingredients Commonly Used for Swine (data on as-fed basis) a

Entry Number Description 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Alfalfa meal dehydrated, 17% CP meal dehydrated, 20% CP Bakery Waste dried bakery product Barley grain, two row grain, six row grain, hulless Beet, Sugar pulp, dried Blood meal, conventional meal, flash dried meal, spray or ring dried plasma, spray dried cells, spray dried Brewers’ Grains dried Buckwheat, Common grain Canola (Rapeseed) meal, sol. extr. Casein dried Cassava (Tapioca or Manioc) meal Coconut (Copra) meal, sol. extr. Corn, Yellow distillers’ grain distillers’ grain with solubles distillers’ solubles, gluten feed gluten meal, 60% CP grain grits by-product (Hominy Feed) Cottonseed meal, mech. extr. 41% CP meal, sol. extr. 41% CP Fababean (Broadbean) seeds Feather meal, hydrolyzed Fish Anchovy meal, mech. extr. Herring meal, mech. extr. Menhaden meal, mech. extr. White meal, mech. extr. solubles, condensed solubles, dried Flax (Linseed) meal sol. extr. Lentil seeds Lupin (Sweet White) seeds Meat meal rendered meal rendered with bone Milk (Cattle) skim, dried

International Feed Number b

Dry Matter (%)

Crude His- IsoProAgri- tileu- Leutein nine dine cine cine (%) (%) (%) (%) (%)

1-00-023 1-00-024

92 92

17.0 19.6

0.71 0.91

0.37 0.68 0.38 0.89

1.21 0.74 0.25 1.40 0.90 0.34

0.18 0.84 0.26 0.93

0.55 0.60

0.70 0.82

0.24 0.35

0.86 1.05

4-00-466

91

10.8

0.46

0.24 0.38

0.80 0.27 0.18

0.23 0.50

0.36

0.33

0.10

0.46

4-00-572 4-00-574 4-00-552

89 89 88

11.3 10.5 14.9

0.54 0.48 0.56

0.25 0.39 0.22 0.37 0.23 0.41

0.77 0.41 0.20 0.68 0.36 0.17 0.77 0.44 0.16

0.28 0.55 0.20 0.49 0.24 0.61

0.29 0.32 0.40

0.35 0.34 0.40

0.11 0.13 0.13

0.52 0.49 0.55

4-00-669

91

8.6

0.32

0.23 0.31

0.53 0.52 0.07

0.06 0.30

0.40

0.38

0.10

0.45

5-00-380 5-26-006 5-00-381 — —

92 92 93 92 92

77.1 87.6 88.8 78.0 92.0

3.34 3.37 3.69 4.55 3.77

5.06 4.57 5.30 2.55 6.99

1.09 1.20 1.04 2.63 0.61

5.34 6.41 5.81 4.42 6.69

2.29 2.32 2.71 3.53 2.14

4.05 4.07 3.78 4.72 3.38

1.08 1.06 1.48 1.36 1.37

7.05 8.03 7.03 4.94 8.50

5-02-141

92

26.5

1.53

0.53 1.02

2.08 1.08 0.45

0.49 1.22

0.88

0.95

0.26

1.26

4-00-994

88

11.1

0.92

0.25 0.40

0.64 0.57 0.19

0.23 0.45

0.31

0.41

0.17

0.56

5-06-145

90

35.6

2.21

0.96 1.43

2.58 2.08 0.74

0.91 1.43

1.13

1.59

0.45

1.82

5-01-162

91

88.7

3.26

2.82 4.66

8.79 7.35 2.70

0.41 4.79

4.77

3.98

1.14

6.10

4-01-152

88

3.3

0.18

0.08 0.11

0.19 0.12 0.04

0.05 0.15

0.04

0.11

0.04

0.14

5-01-573

92

21.9

2.38

0.39 0.75

1.36 0.58 0.35

0.29 0.84

0.58

0.67

0.19

1.07

5-02-842 5-02-843 5-02-844 5-02-903 5-28-242 4-02-935 4-03-011

94 93 92 90 90 89 90

24.8 27.7 26.7 21.5 60.2 8.3 10.3

0.90 1.13 0.90 1.04 1.93 0.37 0.56

0.63 0.69 0.66 0.67 1.28 0.23 0.28

0.28 0.52 0.46 0.46 1.09 0.19 0.18

0.99 1.34 1.38 0.76 3.84 0.39 0.43

0.82 0.83 0.80 0.58 3.25 0.25 0.40

0.62 0.94 1.03 0.74 2.08 0.29 0.40

0.20 0.25 0.23 0.07 0.31 0.06 0.10

1.24 1.30 1.50 1.01 2.79 0.39 0.52

5-01-617 5-07-872

92 90

42.4 41.4

4.26 4.55

1.11 1.29 1.17 1.30

2.45 1.65 0.67 2.47 1.72 0.67

0.69 1.97 0.70 2.20

1.23 1.22

1.34 1.36

0.54 0.48

1.76 1.78

5-09-262

87

25.4

2.28

0.67 1.03

1.89 1.62 0.20

0.32 1.03

0.87

0.89

0.22

1.14

5-03-795

93

84.5

5.62

0.93 3.86

6.79 2.08 0.61

4.13 4.01

2.41

3.82

0.54

5.88

5-01-985 5-02-000 5-02-009 5-02-025 5-01-969 5-01-971

92 93 92 91 51 92

64.6 68.1 62.9 63.3 32.7 64.2

3.68 4.01 3.66 4.04 1.61 2.67

1.56 1.52 1.78 1.34 1.56 1.23

5.00 5.20 4.54 4.39 1.86 2.68

0.61 0.66 0.57 0.68 0.30 0.49

2.66 2.75 2.51 2.32 0.93 1.22

2.15 2.18 2.04 2.03 0.40 0.62

2.82 3.02 2.64 2.60 0.86 1.40

0.76 0.74 0.66 0.66 0.31 0.34

3.51 3.46 3.03 3.06 1.16 1.94

5-02-048

90

33.6

2.97

0.68 1.56

2.06 1.24 0.59

0.59 1.57

1.03

1.26

0.52

1.74

5-02-506

89

24.4

2.05

0.78 1.00

1.84 1.71 0.18

0.27 1.29

0.70

0.84

0.21

1.27

5-27-717

89

34.9

3.38

0.77 1.40

2.43 1.54 0.27

0.51 1.22

1.35

1.20

0.26

1.29

5-00-385 5-00-388

94 93

54.0 51.5

3.60 3.45

1.14 1.60 0.91 1.34

3.84 3.07 0.80 2.98 2.51 0.68

0.60 2.17 0.50 1.62

1.40 1.07

1.97 1.59

0.35 0.28

2.66 2.04

5-01-175

96

34.6

1.24

1.05 1.87

3.67 2.86 0.92

0.30 1.78

1.87

1.62

0.51

0.91 0.88 1.03 2.71 0.49

10.99 11.48 10.81 7.61 12.70

Ly- MePhenylTrypsine thiCys- alaTyro- Thre- to(%) onine tine nine sine onine phan Valine (%) (%) (%) (%) (%) (%) (%)

7.04 7.56 7.45 6.84 8.51

0.99 0.95 0.99 0.75 0.81

0.95 2.63 0.74 0.43 1.03 2.57 0.62 0.50 1.21 2.25 0.82 0.51 0.66 1.96 0.63 0.35 2.48 10.19 1.02 1.43 0.28 0.99 0.26 0.17 0.36 0.98 0.38 0.18

3.06 2.91 2.57 2.61 1.06 1.56

5.11 5.46 4.81 4.51 1.73 2.84

1.95 2.04 1.77 1.76 0.50 0.98

2.33 Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-4

(continued)

Entry Number Description 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79

133

Millet (Proso) grain Oat grain grain, naked groat Pea seeds Peanut (Groundnut) meal, mech. extr. meal, sol. extr. Potato protein concentrate Poultry by-product, meal rendered Rice bran grain, polished ` broken (Brewers’ Rice) polishings Rye grain Safflower meal, sol. extr. meal without hulls, sol. extr. Sesame meal, mech. extr. Sorghum grain Soybean meal, sol. extr. meal without hulls protein concentrate protein isolate seeds, heat processed Sunflower meal, sol. extr. meal without hulls, sol. extr. Triticale grain Wheat bran grain, hard red spring grain, hard red winter grain, soft red winter grain, soft white winter middlings, , 9.5% fiber red dog, , 4% fiber shorts, , 7% fiber Whey dried low lactose, dried permeate, dried Yeast, Brewers’ dehydrated Yeast, Torula dehydrated

International Feed Number b

Dry Matter (%)

Crude His- IsoProAgri- tileu- Leutein nine dine cine cine (%) (%) (%) (%) (%)

4-03-120

90

11.1

0.41

0.20 0.46

1.24 0.23 0.31

0.18 0.56

0.31

0.40

0.16

0.57

4-03-309 4-25-101 4-03-331

89 86 90

11.5 17.1 13.9

0.87 0.77 0.85

0.31 0.48 0.26 0.48 0.24 0.55

0.92 0.40 0.22 0.86 0.47 0.19 0.98 0.48 0.20

0.36 0.65 0.32 0.60 0.22 0.66

0.41 0.42 0.51

0.44 0.40 0.44

0.14 0.16 0.18

0.66 0.63 0.72

5-03-600

89

22.8

1.87

0.54 0.86

1.51 1.50 0.21

0.31 0.98

0.71

0.78

0.19

0.98

5-03-649 5-03-650

92 92

43.2 49.1

4.79 5.09

1.01 1.41 1.06 1.78

2.77 1.48 0.50 2.83 1.66 0.52

0.60 2.02 0.69 2.35

1.74 1.80

1.16 1.27

0.41 0.48

1.70 1.98

5-25-392

91

73.8

3.80

1.71 4.09

7.61 5.83 1.68

1.20 4.89

4.27

4.30

1.02

4.89

5-03-798

93

64.1

3.94

1.25 2.01

3.89 3.32 1.11

0.65 2.26

1.56

2.18

0.48

2.51

4-03-928

90

13.3

1.00

0.34 0.44

0.92 0.57 0.26

0.27 0.56

0.40

0.48

0.14

0.68

4-03-932 4-03-943

89 90

7.9 13.0

0.52 0.82

0.18 0.34 0.28 0.43

0.67 0.30 0.18 0.82 0.58 0.23

0.11 0.39 0.22 0.49

0.38 0.44

0.26 0.44

0.10 0.13

0.49 0.75

4-04-047

88

11.8

0.50

0.24 0.37

0.64 0.38 0.17

0.19 0.50

0.26

0.32

0.12

0.51

5-04-110 5-07-959

92 92

23.4 42.5

2.04 3.59

0.59 0.67 1.07 1.69

1.52 0.74 0.34 2.57 1.17 0.66

0.38 1.07 0.69 2.00

0.77 1.08

0.65 1.28

0.33 0.54

1.18 2.33

5-04-220

93

42.6

4.86

0.98 1.47

2.74 1.01 1.15

0.82 1.77

1.52

1.44

0.54

1.85

4-20-893

88

9.2

0.38

0.23 0.37

1.21 0.22 0.17

0.17 0.49

0.35

0.31

0.10

0.46

5-04-604 5-04-612 — 5-08-038 5-04-597

89 90 90 92 90

43.8 47.5 64.0 85.8 35.2

3.23 3.48 5.79 6.87 2.60

1.17 1.28 1.80 2.25 0.96

3.42 3.66 5.30 6.64 2.75

0.70 0.74 1.00 1.19 0.55

2.18 2.39 3.40 4.34 1.83

1.69 1.82 2.50 3.10 1.32

1.73 1.85 2.80 3.17 1.41

0.61 0.65 0.90 1.08 0.48

2.06 2.27 3.40 4.21 1.68

5-09-340 5-04-739

90 93

26.8 42.2

2.38 2.93

0.66 1.29 0.92 1.44

1.86 1.01 0.59 2.31 1.20 0.82

0.48 1.23 0.66 1.66

0.76 1.03

1.04 1.33

0.38 0.44

1.49 1.74

4-20-362

90

12.5

0.57

0.26 0.39

0.76 0.39 0.20

0.26 0.49

0.32

0.36

0.14

0.51

4-05-190 4-05-258 4-05-268 4-05-294 4-05-337 4-05-205 4-05-203 4-05-201

89 88 88 88 89 89 88 88

15.7 14.1 13.5 11.5 11.8 15.9 15.3 16.0

1.07 0.67 0.60 0.50 0.55 0.97 0.96 1.07

0.44 0.34 0.32 0.20 0.27 0.44 0.41 0.43

0.98 0.93 0.86 0.90 0.79 1.06 1.06 1.02

0.33 0.30 0.29 0.27 0.28 0.32 0.37 0.28

0.62 0.67 0.60 0.63 0.55 0.70 0.66 0.70

0.43 0.40 0.38 0.37 0.36 0.29 0.46 0.51

0.52 0.41 0.37 0.39 0.35 0.51 0.50 0.57

0.22 0.16 0.15 0.26 0.15 0.20 0.10 0.22

0.72 0.61 0.54 0.57 0.53 0.75 0.72 0.87

4-01-182 4-01-186 —

96 96 96

12.1 17.6 3.8

0.26 0.53 0.06

0.23 0.62 0.33 1.16 0.05 0.17

1.08 0.90 0.17 1.61 1.51 0.39 0.22 0.18 0.03

0.25 0.36 0.46 0.63 0.04 0.06

0.25 0.52 —

0.72 1.17 0.14

0.18 0.31 0.03

0.60 1.15 0.13

7-05-527

93

45.9

2.20

1.09 2.15

3.13 3.22 0.74

0.50 1.83

1.55

2.20

0.56

2.39

7-05-534

93

46.4

2.48

1.09 2.50

3.32 3.47 0.69

0.50 2.33

1.65

2.30

0.51

2.60

1.99 2.16 3.30 4.25 1.61

0.49 0.47 0.41 0.45 0.44 0.53 0.55 0.58

MePhenylTrypLy- thiCys- alaTyro- Thre- tosine onine tine nine sine onine phan Valine (%) (%) (%) (%) (%) (%) (%) (%)

2.83 3.02 4.20 5.26 2.22

0.64 0.38 0.34 0.38 0.33 0.57 0.59 0.70

0.61 0.67 0.90 1.01 0.53

0.25 0.23 0.20 0.22 0.20 0.26 0.23 0.25

a

Dash indicates that no data were available. b First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other five digits are the International Feed Number.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

134

Nutrient Requirements of Swine

TABLE 11-5 Entry Number 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

Apparent Ileal Digestibilities of Amino Acids in Some Feed Ingredients Commonly Used for Swinea

Description Alfalfa meal dehydrated, 17% CP meal dehydrated, 20% CP Bakery Waste dried bakery product Barley grain, two row grain, six row grain, hulless Beet, Sugar pulp, dried Blood meal, conventional meal, flash dried meal, spray or ring dried plasma, spray dried cells, spray dried Brewers’ Grains dried Buckwheat, Common grain Canola meal, sol. extr. Casein dried Cassava (Tapioca or Manioc) meal Coconut (Copra) meal, sol. extr. Corn, Yellow distillers’ grain distillers’ grain with solubles distillers’ solubles gluten feed gluten meal, 60% CP grain grits by-product (Hominy Feed) Cottonseed meal, mech. extr. 41% CP meal, sol. extr. 41% CP Fababean (Broadbean) seeds Feather meal, hydrolyzed Fish Anchovy meal, mech. extr. Herring meal, mech. extr. Menhaden meal, mech. extr. White meal, mech. extr. solubles, condensed solubles, dried Flax (Linseed) meal sol. extr. Lentil seeds Lupin (Sweet White) seeds Meat meal rendered meal rendered with bone Milk (Cattle) skim, dried

International Feed Number b

Dry Matter (%)

Crude His- IsoMePhenylTrypProAgri- tileu- Leu- Ly- thiCys- alaTyro- Thre- totein nine dine cine cine sine onine tine nine sine onine phan Valine (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%)

1-00-023 1-00-024

92 92

17.0 19.6

64 —

50 —

59 —

63 —

50 —

64 —

20 —

62 —

59 —

51 —

39 —

55 —

4-00-466

91

10.8

84

84

84

84

62

84

87

88



72

77

81

4-00-572 4-00-574 4-00-552

89 89 88

11.3 10.5 14.9

78 — —

77 — —

75 — —

78 — —

68 — —

80 — —

76 — —

80 — —

78 — —

66 — —

70 — —

73 — —

4-00-669

91

8.6

























5-00-380 5-26-006 5-00-381 — —

92 92 93 91 92

77.1 87.6 88.8 78.0 92.0

56 — 91 90 —

60 — 92 91 —

55 — 71 85 —

60 — 91 84 —

56 — 91 87 —

42 — 85 64 —

55 — 81 — —

60 — 90 88 —

— — 88 — —

54 — 86 82 —

65 — 88 92 —

54 — 90 86 —

5-02-141

92

26.5

81

70

81

73

69

74

67

81

91

70

73

73

4-00-994

88

11.1

























5-06-145

90

35.6

81

80

74

78

74

82

79

76

73

69

73

71

5-01-162

91

88.7

94

95

92

96

95

96

77

95

96

88

92

94

4-01-152

88

3.3

























5-01-573

92

21.9

81

63

64

68

51

67

54

71



51

63

68

5-02-842 5-02-843 5-02-844 5-02-903 5-28-242 4-02-935 4-03-011

94 93 92 90 90 89 90

24.8 27.7 26.7 21.5 60.2 8.3 10.3

— 72 — 79 87 83 —

— 61 — 69 82 82 —

— 66 — 68 84 79 —

— 76 — 81 88 88 —

— 47 — 51 75 66 —

— 72 — 79 87 86 —

— 57 — 53 79 78 —

— 76 — 80 86 83 —

— 71 — 80 84 83 —

— 55 — 57 80 69 —

— 50 — 47 81 64 —

— 63 — 71 82 79 —

5-01-617 5-07-872

92 90

42.4 41.4

— 88

— 77

— 69

— 70

— 61

— 73

— 68

— 81

— 77

— 63

— 67

— 71

5-09-262

87

25.4

89

85

80

82

84

73

65

78

79

75

68

78

5-03-795

93

84.5

81

56

81

80

54

65

71

82

73

74

63

80

5-01-985 5-02-000 5-02-009 5-02-025 5-01-969 5-01-971

92 93 92 91 51 92

64.6 68.1 62.9 63.3 32.7 64.2

— — 90 — — —

— — 86 — 90 —

— — 87 — 88 —

— — 88 — 91 —

— — 89 — 92 —

— — 88 — 92 —

— — 73 — 61 —

— — 85 — 91 —

— — 86 — 82 —

— — 85 — 88 —

— — 79 — 63 —

— — 85 — 88 —

5-02-048

90

33.6

86

72

75

68

70

76



78



63

75

74

5-02-506

89

24.4

81

76

75

76

83

79



71



70



72

5-27-717

89

34.9

91

85

82

81

78

65

78

82

81

74



77

5-00-385 5-00-388

94 93

54.0 51.5

88 81

82 75

82 74

82 76

83 74

85 79

55 55

83 76

79 71

79 70

73 60

79 74

5-01-175

96

34.6

89

93

86

93

91

92

81

93

94

85

90

87 Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-5 Entry Number 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79

135

(continued)

Description Millet (Proso) grain Oat grain grain, naked groat Pea seeds Peanut (Groundnut) meal, mech. extr. meal, sol. extr. Potato protein concentrate Poultry by-product, meal rendered Rice bran grain, polished ` broken (Brewers’ Rice) polishings Rye grain Safflower meal, sol. extr. meal without hulls, sol. extr. Sesame meal, mech. extr. Sorghum grain Soybean meal, sol. extr. meal without hulls protein concentrate protein isolate seeds, heat processed Sunflower meal, sol. extr. meal without hulls, sol. extr. Triticale grain Wheat bran grain, hard red spring grain, hard red winter grain, soft red winter grain, soft white winter middlings, , 9.5 % fiber red dog, , 4% fiber shorts, , 7% fiber Whey dried low lactose, dried permeate, dried Yeast, Brewers’ dehydrated Yeast, Torula dehydrated

International Feed Number b

Dry Matter (%)

Crude His- IsoMePhenylTrypProAgri- tileu- Leu- Ly- thiCys- alaTyro- Thre- totein nine dine cine cine sine onine tine nine sine onine phan Valine (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%)

4-03-120

90

11.1

82

85

83

87

74

72

82

85



75

84

81

4-03-309 4-25-101 4-03-331

89 86 90

10.8 17.1 13.9

85 — 86

81 — 83

74 — 83

78 — 83

70 — 79

79 — 85

69 — 80

81 — 86

76 — 82

59 — 76

72 — 80

73 — 82

5-03-600

89

22.8

87

83

79

80

84

78

68

81

83

73

70

76

5-03-649 5-03-650

92 92

43.2 49.1

— 93

— 81

— 83

— 85

— 78

— 85

— 77

— 89

— 91

— 74

— 73

— 82

5-25-392

91

73.8

83

84

80

83

79

83

56

82

78

78

59

78

5-03-798

93

64.1

85

76

77

78

78

74

70

80

71

72

74

74

4-03-928

91

13.3

85

78

64

65

72

74

66

68

77

61

64

66

4-03-932 4-03-943

89 90

7.9 13.0

— 82

— 80

— 62

— 65

— 68

— 71

— 61

— 64

— 68

— 61

— 61

— 63

4-04-047

88

11.8

73

71

68

71

64

76

74

76

65

59

67

67

5-04-110 5-07-959

92 92

23.4 42.5

— —

— —

— —

— —

— —

— —

— —

— —

— —

— —

— —

— —

5-04-220

93

42.6

94

76

85

85

76

90

86

89

87

78

85

84

4-20-893

88

9.2

78

73

80

86

62

81

79

81

83

68

75

78

5-04-604 5-04-612 — 5-08-038 5-04-597

89 90 90 92 90

43.8 47.5 64.0 85.8 35.2

91 90 97 91 89

86 86 95 88 82

84 84 93 90 78

84 84 93 92 80

85 85 93 88 81

86 86 91 — 78

77 79 90 — 76

85 84 94 88 82

86 85 93 89 85

78 78 90 85 77

80 81 89 — 75

81 81 91 86 76

5-09-340 5-04-739

90 93

26.8 42.2

90 89

79 79

79 78

79 77

75 74

88 87

75 74

82 80

83 77

74 71

79 76

77 75

4-20-362

90

12.5

85

84

80

82

76

85

83

84

81

69

74

79

4-05-190 4-05-258 4-05-268 4-05-294 4-05-337 4-05-205 4-05-203 4-05-201

89 88 88 88 89 89 88 88

15.7 14.1 13.5 11.5 11.8 15.9 15.3 16.0

83 — — 83 — 88 — 86

76 — — 84 — 76 — 82

69 — — 84 — 77 — 77

71 — — 85 — 78 — 80

69 — — 73 — 75 — 73

76 — — 85 — 82 — 81

70 — — 84 — 82 — 66

76 — — 87 — 83 — 82

75 — — 84 — 83 — 78

60 — — 72 — 69 — 72

65 — — 81 — 77 — 77

70 — — 80 — 76 — 76

4-01-182 4-01-186 —

96 96 96

12.1 17.6 3.8

86 83 —

91 92 —

85 90 —

89 94 —

82 85 —

84 92 —

86 92 —

80 92 —

71 92 —

79 87 —

78 92 —

81 89 —

7-05-527

93

45.9

79

77

74

73

76

72

38

72

61

63

60

70

7-05-534

93

46.4

























a

Dash indicates that no data were available. b First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplement; 6, minerals; 7, vitamins; 8, additives; the otherˆ five digits are the International Feed Number. Source: Southern (1991), Rhone-Poulenc (1993a), Jondreville et al. (1995), and Heartland Lysine (1995).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

136

Nutrient Requirements of Swine

TABLE 11-6 Entry Number 01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41

True Ileal Digestibility of Amino Acids in Some Feed Ingredients Commonly Used for Swine a

Description Alfalfa meal dehydrated, 17% CP meal dehydrated, 20% CP Bakery Waste dried bakery product Barley grain, two row grain, six row grain, hulless Beet, Sugar pulp, dried Blood meal, conventional meal, flash dried meal, spray or ring dried plasma, spray dried cells, spray dried Brewers’ Grains dried Buckwheat, Common grain Canola meal, sol. extr. Casein dried Cassava (Tapioca or Manioc) meal Coconut (Copra) meal, sol. extr. Corn, Yellow distillers’ grain distillers’ grain with solubles distillers’ solubles gluten feed gluten meal, 60% CP grain grits by-product (Hominy Feed) Cottonseed meal, mech. extr. 41% CP meal, sol. extr. 41% CP Fababean (Broadbean) seeds Feather meal, hydrolyzed Fish Anchovy meal, mech. extr. Herring meal, mech. extr. Menhaden meal, mech. extr. White meal, mech. extr. solubles, condensed solubles, dried Flax (Linseed) meal sol. extr. Lentil seeds Lupin (Sweet White) seeds Meat meal rendered meal rendered with bone Milk (Cattle) skim, dried

International Feed Number b

Dry Matter (%)

Crude His- IsoMePhenylTrypProAgri- tileu- Leu- Ly- thiCys- alaTyro- Thre- totein nine dine cine cine sine onine tine nine sine onine phan Valine (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%)

1-00-023 1-00-024

92 92

17.0 19.6

74 —

59 —

68 —

71 —

56 —

71 —

37 —

70 —

66 —

63 —

46 —

64 —

4-00-466

91

10.8

























4-00-572 4-00-574 4-00-552

89 89 88

11.3 10.5 14.9

86 — —

86 — —

84 — —

86 — —

79 — —

86 — —

86 — —

88 — —

87 — —

81 — —

80 — —

82 — —

4-00-669

91

8.6

57

61

60

59

51

64

21

54

51

30

41

42

5-00-380 5-26-006 5-00-381 — —

92 92 93 91 92

77.1 87.6 88.8 78.0 92.0

— — 92 — —

— — 92 — —

— — 88 — —

— — 92 — —

— — 94 — —

— — 96 — —

— — 91 — —

— — 93 — —

— — 93 — —

— — 94 — —

— — 94 — —

— — 91 — —

5-02-141

92

26.5

95

84

89

88

82

90

77

92

94

81

83

86

4-00-994

88

11.1

























5-06-145

90

35.6

85

85

78

81

78

86

83

82

79

76

75

77

5-01-162

91

88.7

























4-01-152

88

3.3

91

76

29

75

64

82

62

76

66

69



74

5-01-573

92

21.9

























5-02-842 5-02-843 5-02-844 5-02-903 5-28-242 4-02-935 4-03-011

94 93 92 90 90 89 90

24.8 27.7 26.7 21.5 60.2 8.3 10.3

— 77 — 87 89 89 —

— 61 — 78 80 87 —

— 73 — 80 84 87 —

— 79 — 85 88 92 —

— 59 — 66 80 78 —

— 75 — 83 90 90 —

— 60 — 59 82 86 —

— 79 — 87 85 90 —

— 77 — 84 87 89 —

— 65 — 71 84 82 —

— — — 64 63 84 —

— 67 — 77 80 87 —

5-01-617 5-07-872

92 90

42.4 41.4

— 89

— 79

— 71

— 73

— 64

— 75

— 69

— 81

— 78

— 68

— 65

— 72

5-09-262

87

25.4

91

87

84

86

87

81

76

85

82

82

75

82

5-03-795

93

84.5

85

74

88

84

67

74

73

86

79

82

86

84

5-01-985 5-02-000 5-02-009 5-02-025 5-01-969 5-01-971

92 93 92 91 51 92

64.6 68.1 62.9 63.3 32.7 64.2

— — 94 — 98 —

— — 93 — 93 —

— — 94 — 94 —

— — 94 — 96 —

— — 95 — 95 —

— — 94 — 95 —

— — 88 — 78 —

— — 93 — 96 —

— — 92 — 94 —

— — 88 — 95 —

— — 90 — 91 —

— — 93 — 94 —

5-02-048

90

33.6

























5-02-506

89

24.4

























5-27-717

89

34.9

92

88

83

83

79

68

84

85

85

79



80

5-00-385 5-00-388

94 93

54.0 51.5

86 83

83 83

84 82

83 81

83 80

87 83

58 63

85 81

80 78

82 80

79 78

80 79

5-01-175

96

34.6

92

96

88

97

93

96

89

98

97

92

97

91 Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-6 Entry Number 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79

137

(continued)

Description Millet (Proso) grain Oat grain grain, naked groat Pea seeds Peanut (Groundnut) meal, mech. extr. meal, sol. extr. Potato protein concentrate Poultry by-product, meal rendered Rice bran grain, polished ` broken (Brewers’ Rice) polishings Rye grain Safflower meal, sol. extr. meal, without hulls, sol. extr. Sesame meal, mech. extr. Sorghum grain Soybean meal, sol. extr. meal without hulls protein concentrate protein isolate seeds, heat processed Sunflower meal, sol. extr. meal without hulls, sol. extr. Triticale grain Wheat bran grain, hard red spring grain, hard red winter grain, soft red winter grain, soft white winter middlings, , 9.5% fiber red dog, , 4% fiber shorts, , 7% fiber Whey dried low lactose, dried permeate, dried Yeast, Brewers’ dehydrated Yeast, Torula dehydrated

International Feed Number b

Dry Matter (%)

Crude His- IsoMePhenylTrypProAgri- tileu- Leu- Ly- thiCys- alaTyro- Thre- totein nine dine cine cine sine onine tine nine sine onine phan Valine (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%) (%)

4-03-120

90

11.1

























4-03-309 4-25-101 4-03-331

89 86 90

10.8 17.1 13.9

89 — 86

85 — 82

80 — 83

83 — 83

76 — 79

84 — 86

75 — 85

86 — 84

82 — 84

71 — 80

78 — 82

79 — 81

5-03-600

89

22.8

90

89

85

86

88

84

79

87

87

83

81

83

5-03-649 5-03-650

92 92

43.2 49.1

— 97

— 91

— 92

— 93

— 88

— 89

— 86

— 94

— 95

— 90

— —

— 91

5-25-392

91

73.8

88

86

82

85

81

86

64

85

80

83

66

81

5-03-798

93

64.1

85

78

81

80

80

77

72

81

76

77



74

4-03-928 4-03-932

91 89

13.3 7.9

89 —

87 —

69 —

70 —

78 —

77 —

68 —

73 —

81 —

71 —

— —

69 —

4-03-943

90

13.0

85

82

67

69

72

75

65

68

73

67

67

67

4-04-047

88

11.8

79

78

77

79

73

81

83

82

76

73

75

75

5-04-110 5-07-959

92 92

23.4 42.5

— —

— —

— —

— —

— —

— —

— —

— —

— —

— —

— —

— —

5-04-220

93

42.6

96

93

91

92

85

92

92

93

91

90



89

4-20-893

88

9.2

87

81

87

90

81

89

83

88

87

84

83

87

5-04-604 5-04-612 — 5-08-038 5-04-597

89 90 90 92 90

43.8 47.5 64.0 85.8 35.2

93 94 99 — 93

90 91 97 — 88

88 89 95 — 84

88 89 95 — 86

89 90 95 — 86

91 91 94 — 85

84 87 94 — 80

88 89 97 — 88

90 90 96 — 87

85 87 94 — 83

87 90 93 — 82

86 88 94 — 83

5-09-340 5-04-739

90 93

26.8 42.2

93 93

83 85

84 84

85 85

81 83

91 90

81 81

87 86

88 88

82 84

84 —

82 82

4-20-362

90

12.5

88

84

84

86

81

89

87

85

83

76

88

84

4-05-190 4-05-258 4-05-268 4-05-294 4-05-337 4-05-205 4-05-203 4-05-201

89 88 88 88 89 89 88 88

15.7 14.1 13.5 11.5 11.8 15.9 15.3 16.0

87 — — 88 — 95 — 89

82 — — 89 — 94 — 84

76 — — 89 — 92 — 81

78 — — 89 — 93 — 84

71 — — 81 — 89 — 77

79 — — 90 — 93 — 85

77 — — 90 — 91 — 80

81 — — 91 — 95 — 86

80 — — 89 — 92 — 83

70 — — 84 — 88 — 78

74 — — 90 — 91 — 83

75 — — 86 — 90 — 81

4-01-182 4-01-186 —

96 96 96

12.1 17.6 3.8

48 88 —

89 95 —

83 92 —

87 96 —

87 87 —

81 95 —

85 96 —

83 96 —

77 95 —

79 89 —

79 95 —

77 92 —

7-05-527

93

45.9

78

77

72

73

74

71

48

67

64

66

54

66

7-05-534

93

46.4

























a

Dash indicates that no data were available. b First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other ˆ five digits are the International Feed Number. Source: Southern (1991), Rhone-Poulenc (1993a), and Jondreville et al. (1995).

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

138

Nutrient Requirements of Swine

TABLE 11-7

Coefficients for Estimation of Amino Acids from Crude Protein Content of Feed Ingredients a,b

Ingredient

Dry Matter (%)

Crude Protein (%)

Regression Factors

Alfalfa meal

88

17.0

Bakery waste

88

10.6

Barley

88

10.6

Brewers’ grains

88

22.8

Canola meal (Rapeseed)

88

34.8

a b r a b r a b r a b r a b

Coconut meal

88

18.6

Corn

88

8.5

Corn distillers’ grains with solubles

88

27.7

Corn gluten feed

88

18.9

Corn gluten meal

88

60.6

Cottonseed meal

88

41.9

Fababean seeds

88

25.4

Fish meal

91

62.9

Lupin seeds

88

33.8

Meat and bone meal

91

49.1

Meat meal

91

48.8

Milk, dried skim

93

35.8

Oats

88

12.6

Peanut meal

88

43.2

Peas, seeds

88

20.9

Potato protein concentrate

88

73.8

Poultry by-product meal

91

57.7

Rice bran

88

13.1

Rye

88

9.6

a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r a b r

Lysine

Tryptophan

10.2140 0.0561 0.86 10.0310 0.0284 0.90 0.1330 0.0235 0.83 0.1800 0.0295 0.73 0.0520 0.0547 0.53 0.1500 0.0174 0.74 0.0790 0.0186 0.62 0.0090 0.0221 0.94 10.2440 0.0433 0.64 — — — 10.1250 0.0440 0.82 0.1120 0.0598 0.78 11.9980 0.1081 0.86 0.5510 0.0294 0.86 11.0560 0.0729 0.82 10.8780 0.0694 0.80 10.4360 0.0893 0.75 0.0780 0.0358 0.94 0.2300 0.0290 0.76 0.4830 0.0485 0.75 11.2540 0.0960 0.50 10.2600 0.0620 0.72 0.0220 0.0446 0.96 — — —

10.0350 0.0160 0.89 0.0110 0.0093 0.98 0.0230 0.0095 0.88 0.0690 0.0086 0.90 10.1750 0.0181 0.71 10.0040 0.0080 0.98 0.0210 0.0047 0.65 — — — — — — 10.0660 0.0063 0.59 10.0510 0.0132 0.92 0.0540 0.0109 0.71 10.3880 0.0158 0.76 0.0230 0.0069 0.91 10.4030 0.0139 0.76 10.3150 0.0127 0.74 10.2320 0.0102 0.92 10.0170 0.0135 0.92 10.0277 0.0164 0.90 0.0500 0.0066 0.64 10.6410 0.0226 0.63 10.2830 0.0135 0.71 10.0800 0.0182 0.97 0.0420 0.0054 0.61

Theonine

Methionine

Methionine ` Cystine

10.0850 0.0460 0.89 10.0150 0.0311 0.97 0.0440 0.0299 0.96 0.0730 0.0333 0.98 0.4800 0.0303 0.63 0.0200 0.0297 0.92 0.0300 0.0326 0.93 0.6150 0.0118 0.70 10.1340 0.0430 0.88 0.3030 0.0293 0.76 0.1530 0.0289 0.88 0.1920 0.0278 0.88 10.7420 0.0537 0.85 0.3550 0.0250 0.93 10.8060 0.0488 0.86 10.5460 0.0447 0.86 0.3720 0.0337 0.67 0.0210 0.0329 0.98 0.3780 0.0181 0.93 0.3490 0.0207 0.72 11.7150 0.0815 0.60 10.7270 0.0504 0.79 0.0310 0.0359 0.95 0.0740 0.0281 0.67

10.0720 0.0188 0.92 10.0310 0.0179 0.94 0.0190 0.0152 0.92 10.1270 0.0250 0.95 0.1410 0.0164 0.65 10.0460 0.0175 0.86 0.0330 0.0170 0.70 0.2870 0.0076 0.73 10.0310 0.0184 0.68 — — — 0.1070 0.0135 0.80 0.0210 0.0072 0.63 10.6900 0.0391 0.82 10.2020 0.0138 0.91 10.4390 0.0228 0.74 10.2210 0.0184 0.80 0.1150 0.0216 0.65 10.0140 0.0182 0.96 0.1290 0.0087 0.64 — — — 0.3050 0.0186 0.50 10.4940 0.0278 0.75 10.0400 0.0240 0.94 10.0130 0.0194 0.76

0.0240 0.0267 0.92 `0.0500 0.0404 0.97 0.1010 0.0301 0.89 10.0580 0.0436 0.93 10.0310 0.0467 0.72 10.0700 0.0349 0.95 0.1290 0.0283 0.72 — — — — — — — — — 10.0780 0.0347 0.83 0.1290 0.0155 0.68 10.5710 0.0463 0.78 10.2470 0.0303 0.87 10.7240 0.0387 0.70 10.5480 0.0366 0.66 0.2720 0.0252 0.64 0.0390 0.0424 0.96 0.1540 0.0219 0.78 — — — — — — 10.5660 0.0404 0.65 0.0040 0.0425 0.97 10.0240 0.0460 0.77 Continues

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients TABLE 11-7

139

(continued)

Ingredient

Dry Matter (%)

Crude Protein (%)

Regression Factors

Lysine

Tryptophan

Sesame meal

88

41.1

Sorghum grain

88

9.2

Soybean meal

88

45.6

Sunflower meal

88

33.5

Triticale

88

11.6

Wheat

88

13.3

Wheat bran

88

15.7

Wheat middlings

88

15.9

a b r a b r a b r a b r a b r a b r a b r a b r

0.1540 0.0210 0.81 0.0910 0.0138 0.76 10.0810 0.0644 0.78 0.1720 0.0304 0.86 0.2050 0.0183 0.61 10.0270 0.0306 0.77 0.0400 0.0381 0.80 0.3230 0.0158 0.41

10.1680 0.0181 0.94 0.0170 0.0090 0.94 0.0580 0.0118 0.59 10.0490 0.0134 0.92 0.0260 0.0081 0.83 0.0310 0.0091 0.85 0.0650 0.0099 0.50 — — —

Theonine 0.1760 0.0308 0.94 0.0320 0.0302 0.98 0.0810 0.0381 0.81 0.0360 0.0361 0.95 0.1390 0.0214 0.71 0.0080 0.0284 0.94 0.0470 0.0299 0.89 0.1300 0.0240 0.96

Methionine

Methionine ` Cystine

10.0080 0.0282 0.88 0.0390 0.0140 0.86 0.0170 0.0141 0.65 10.0570 0.0247 0.93 0.0550 0.0131 0.780 0.0030 0.0157 0.92 0.0030 0.0155 0.82 0.0690 0.0123 0.73

0.0250 0.0473 0.91 0.0980 0.0261 0.87 0.1470 0.0263 0.57 10.0160 0.0411 0.94 0.1310 0.0309 0.75 0.0750 0.0322 0.93 0.1620 0.0264 0.80 10.0250 0.0387 0.88

a To estimate amino acid content, use the equation y 4 a ` bx, where y is the percentage of the amino acid in the sample, x is the percentage of crude protein in the sample, a is the intercept, and b is the regression coefficient. The r-value is the correlation between the two variables. Note that the percentages of crude protein and dry matter may not agree with previous tables because they were obtained from different data sets. b Dash indicates that no coefficients were available. Source: Fickler et al. (1995).

TABLE 11-8

Mineral Concentrations in Macro Mineral Sources (data on as-fed basis)

Entry Number Description

International Feed No.

Calcium a (%)

Phosphorus Phos- BioavailChlo- Potas- Magnephorus ability b Sodium rine sium sium Sulfur Iron (%) (%) (%) (%) (%) (%) (%) (%)

Manganese (%)

01 02 03 04 05 06 07 08 09 10 11 12 13 14 15 16 17 18 19 20 21 22

6-00-400 6-01-069 6-01-080 6-26-334 6-01-090 6-02-632 6-02-754 6-02-756 6-02-758 6-01-780 6-09-338 6-05-586 6-03-947 6-03-755 6-06-177 6-08-098 6-12-316 6-04-272 6-04-152 6-04-286 6-04-288 6-04-291

29.80 38.50 20 to 24 17.00 21.85 35.84 0.02 1.69 0.02 32.00 0.35 35.09 16.09 0.05 0.06 0.15 — 0.01 0.30 — 0.09 —

12.50 0.02 18.50 21.10 — 0.01 — — — 18.00 24.20 14.23 9.05 — — — — — — 21.15 24.94 —

0.03 0.02 0.14 0.01 — 0.02 0.01 — — 0.05 0.01 — 0.10 0.001 0.002 0.001 — — — — — —

Bone meal, steamed Calcium carbonate Calcium phosphate (dicalcium) Calcium phosphate (monocalcium) Calcium sulfate, dihydrate Limestone, groundc Magnesium carbonate Magnesium oxide Magnesium sulfate, heptahydrate Phosphate, defluorinated Phosphate, monoammonium ¸ Phosphate, rock curacao, ground Phosphate, rock, soft Potassium chloride Potassium and magnesium sulfate Potassium sulfate Sodium carbonate Sodium bicarbonate Sodium chloride Sodium phosphate, dibasic Sodium phosphate, monobasic Sodium sulfate, decahydrate

80 to 90 95 to 100 100

85 to 95 100 40 to 60 30 to 50

100 100

0.04 0.08 0.18 0.20 — 0.06 — — — 3.27 0.20 0.20 0.10 1.00 0.76 0.09 43.30 27.00 39.50 31.04 18.65 13.80

— 0.02 0.47 — — 0.02 — — 0.01 — — — — 46.93 1.25 1.50 — — 59.00 — 0.02 —

0.20 0.08 0.15 0.16 — 0.11 — 0.02 — 0.10 0.16 — — 51.37 18.45 43.04 — 0.01 — — 0.01 —

0.30 1.61 0.80 0.90 0.48 2.06 30.20 55.00 9.60 0.29 0.75 0.80 0.38 0.23 11.58 0.60 — — 0.005 — 0.01 —

2.40 0.08 0.80 0.80 16.19 0.04 — 0.10 13.04 0.13 1.50 — — 0.32 21.97 17.64 — — 0.20 — — 9.70

— 0.06 0.79 0.75 — 0.35 — 1.06 — 0.84 d 0.41 0.35 1.92 0.06 0.01 0.07 — — 0.01 — — —

NOTE: The mineral supplements used as feed supplements are not chemically pure compounds, and the composition may vary substantially among sources. The supplier’s analysis should be used if it is available. For example, feed-grade dicalcium phosphate contains some monocalcium phosphate and feed-grade monocalcium phosphate contains some dicalcium phosphate. Dashes indicate that no data were available. a Estimates suggest 90 to 100% bioavailability of calcium in most sources of monocalcium phosphate, dicalcium phosphate, tricalcium phosphate, defluorinated phosphate, calcium carbonate, calcium sulfate, and calcitic limestone. The calcium in high-magnesium limestone or dolomitic limestone is less bioavailable (50 to 80%). b Bioavailability estimates are generally expressed as a percentage of monosodium phosphate or monocalcium phosphate. c Most calcitic limestones will contain 38% or more calcium and less magnesium than shown. d Iron in defluorinated phosphate is about 65% as available as the iron in ferrous sulfate.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

140

Nutrient Requirements of Swine Table 11-9

Inorganic Sources and Estimated Bioavailabilities of Trace Minerals a Chemical Formula

Mineral Content (%)

Relative Bioavailability (%)

Copper Cupric sulfate (pentahydrate) Cupric chloride, tribasic Cupric oxide Cupric carbonate (monohydrate) Cupric sulfate (anhydrous)

CuSO4•5H2O Cu2(OH)3Cl CuO CuCO3•Cu(OH)2•H2O CuSO4

25.2 58.0 75.0 50 to 55 39.9

100 100 0 to 10 60 to 100 100

Iron Ferrous sulfate (monohydrate) Ferrous sulfate (heptahydrate) Ferrous carbonate Ferric oxide Ferric chloride (hexahydrate) Ferrous oxide

FeSO4•H2O FeSO4•7H2O FeCO3 Fe2O3 FeCl3•6H2O FeO

30.0 20.0 38.0 69.9 20.7 77.8

100 100 15 to 80 0 40 to 100 —c

Iodine Ethylenediamine dihydroiodide (EDDI) Calcium iodate Potassium iodide Potassium iodate Cupric iodide

C2H8N22HI Ca(IO3)2 KI KIO3 CuI

79.5 63.5 68.8 59.3 66.6

100 100 100 —c 100

Manganese Manganous Manganous Manganous Manganous Manganous

MnSO4•H2O MnO MnO2 MnCO3 MnCl2•4H2O

29.5 60.0 63.1 46.4 27.5

100 70 35 to 95 30 to 100 100

Selenium Sodium selenite Sodium selenate (decahydrate)

Na2SeO3 Na2SeO4•10H2O

45.0 21.4

100 100

Zinc Zinc Zinc Zinc Zinc Zinc

ZnSO4•H2O ZnO ZnSO4•7H2O Zn•CO3 ZnCl2

35.5 72.0 22.3 56.0 48.0

100 50 to 80 100 100 100

Mineral Element and Source

b

sulfate (monohydrate) oxide dioxide carbonate chloride (tetrahydrate)

sulfate (monohydrate) oxide sulfate (heptahydrate) carbonate chloride

a

The mineral source listed first under each mineral element was generally the standard with which the other sources were compared to establish relative bioavailability. b Less commonly used sources in italic. c —indicates no data available.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Composition of Feed Ingredients Table 11-10

141

Characteristics and Energy Values of Various Sources of Fats and Oils (data on as-fed basis) a,b

Type of Lipid

International Feed Number c

Selected Fatty Acids (% of Total Fatty Acids)

Energy Content (kcal/kg)

#C10

C12:0 C14:0 C16:0 C16:1 C18:0 C18:1 C18:2 C18:3 $20

Total sat.

Total unsat.

Animal Fats Beef Tallow Choice White Grease Lard Poultry Fat Restaurant Grease

4-08-127 — 4-04-790 4-09-319 —

0.0 0.2 0.1 0.0 —

0.9 0.2 0.2 0.1 —

2.7 1.9 1.3 0.9 1.9

24.9 21.5 23.8 21.6 16.2

4.2 5.7 2.7 5.7 2.5

18.9 14.9 13.5 6.0 10.5

36.0 41.1 41.2 37.3 47.5

3.1 11.6 10.2 19.5 17.5

0.6 0.4 1.0 1.0 1.9

0.3 1.8 1.0 1.2 1.0

52.1 40.8 41.1 31.2 29.9

47.9 59.2 58.9 68.8 70.1

0.92 1.45 1.44 2.20 2.34

Fish Oils Anchovy Herring Menhaden

— 7-08-048 7-08-049

— — —

— 0.2 —

7.4 7.1 8.0

17.4 11.7 15.1

10.5 9.6 10.5

4.0 0.8 3.8

11.6 11.9 14.5

1.2 1.1 2.1

0.8 0.8 1.5

30.3 45.6 29.5

34.6 22.8 33.3

65.4 77.2 66.7

1.89 3.39 2.00

Vegetable Oils Canola (Rapeseed) Coconut Corn Cottonseed Olive Palm Peanut Safflower Sesame Soybean Sunflower

4-06-144 4-09-320 4-07-882 4-20-836 — — 4-03-658 4-20-526 — 4-07-983 4-20-833

0.0 14.1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

0.0 44.6 0.0 0.0 0.0 0.1 0.0 0.0 0.0 0.0 0.0

0.0 16.8 0.0 0.8 0.0 1.0 0.1 0.1 0.0 0.1 0.0

4.0 8.2 10.9 22.7 11.0 43.5 9.5 6.2 8.9 10.3 5.4

0.2 0.0 0.0 0.8 0.8 0.3 0.1 0.4 0.2 0.2 0.2

1.8 2.8 1.8 2.3 2.2 4.3 2.2 2.3 4.8 3.8 3.5

56.1 5.8 24.2 17.0 72.5 36.6 44.8 11.7 39.3 22.8 45.3

20.3 1.8 59.0 51.5 7.9 9.1 32.0 74.1 41.3 51.0 39.8

9.3 0.0 0.7 0.2 0.6 0.2 — 0.4 0.3 6.8 0.2

3.6 — — 0.1 0.3 0.1 6.4 — 0.2 0.2 —

7.4 91.9 13.3 27.1 14.1 51.6 17.8 9.5 14.8 15.1 10.6

92.6 8.1 86.7 72.9 85.9 48.4 82.2 90.5 85.2 84.9 89.4

12.46 0.09 6.53 2.69 6.08 0.94 4.63 9.52 5.73 5.64 8.47

U:Sd ratio

Iodine value

Total SN-6

Total SN-3

DE d

ME e

NE f

3.1 11.6 10.2 19.5 17.5

0.6 0.4 1.0 1.0 1.9

8,000 8,290 8,285 8,520 8,550

7,680 7,955 7,950 8,180 8,205

4,925 5,095 5,100 5,230 5,245

— — —

1.3 1.4 1.5

31.2 17.8 25.1

8,445 8,680 8,475

8,105 8,330 8,135

5,185 5,320 5,200

118 10 125 105 86 50 92 140 110 130 133

20.3 1.8 58.0 51.5 7.9 9.1 32.0 74.1 41.3 51.0 39.8

9.3 0.0 0.7 0.2 0.6 0.2 0.0 0.4 0.3 6.8 0.2

8,760 8,405g 8,755 8,605 8,750 8,010 8,735 8,760 8,750 8,750 8,760

8,410 8,070 8,405 8,260 8,400 7,690 8,385 8,410 8,400 8,400 8,410

5,365 5,160 5,360 5,275 5,360 4,935 5,350 5,365 5,360 5,360 5,365

44 60 64 78 75

a

Dash indicates that no data were available. The fatty acid data were obtained from Pearl (1995) of the Fats and Protein Research Foundation and USDA Food Composition Standard Release 11 (1997). Values for fatty acid content do not always total 100% but represent means as obtained from various fat analysis conducted by gas-liquid chromatography. c First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other five digits are the International Feed Number. d Calculated by the following relationship (Powles et al. 1995): DE(kcal/kg) 4 (36.898 1 (0.005 2 FFA) 1 (7.330 2 e10.9062U:S))/4.184 where FFA is the free fatty acid content in g/kg and U:S is the ratio of unsaturated to saturated fatty acids. In calculating the DE, the free fatty acid concentrations of all fats were assumed to be 50 g/kg (or 5%). e Calculated as 96% of DE. f Calculated by Equation 1-12 in Chapter 1. g Coconut oil was considered outside the range of the data used to develop the relationship in footnote d. The DE concentration of coconut oil was calculated from the digestibility (89.42% of GE) reported by Cera et. al (1989) for pigs from 2 to 4 weeks after weaning at 3 weeks of age. b

TABLE 11-11 Chemical Composition of Some Purified Feed Ingredients Commonly Used for Swine Research (data on as-fed basis)a Entry Number 01 02 03 04 05 06

Description

International Feed Number b

Dry Matter (%)

DE (kcal/ kg)

ME (kcal/ kg)

NE (kcal/ kg)

Crude Protein (%)

Crude Fat (%)

Linoleic Acid (%)

Calcium (%)

Phosphorus (%)

Lysine c

Casein Corn starch Glucose monohydrate Lactose Gelatin Sucrose

5-01-162 4-02-889 4-02-125 4-07-881 5-14-503 4-04-701

91 99 90 96 90 99

4,135 4,000 3,360 3,525 2,800 3,795

3,535 3,985 3,260 3,435 2,140 3,635

2,555 2,505 1,940 2,370 d 1,570 d 2,730

88.7 0.3 0.3 0.3 88.6 0.0

0.80 0.22 — — 0.50 0.00

0.03 — — — — —

0.63 0.00 — — 0.49 0.04

1.01 0.03 — — — 0.01

7.35 — — — 3.62 —

a

Dash indicates that no data were available. First digit is class of feed: 1, dry forages and roughages; 2, pasture, range plants, and forages fed green; 3, silages; 4, energy feeds; 5, protein supplements; 6, minerals; 7, vitamins; 8, additives; the other five digits are the International Feed Number. c Amino acid composition of casein is shown in Table 11-4. Other amino acids in gelatin: arginine, 6.60%; histidine, 0.66%; isoleucine, 1.42%; leucine, 2.91%; methionine, 0.76%; cystine, 0.12%; phenylalanine, 1.74%; tyrosine, 0.43%; threonine, 1.82%; tryptophan, 0.05%; and valine, 2.26%. d Calculated by Equation 1-12 in Chapter 1. b

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

142

Nutrient Requirements of Swine

REFERENCES Ammerman, C. B., D. H. Baker, and A. J. Lewis. 1995. Bioavailability of Nutrients for Animals. Amino Acids, Minerals and Vitamins. Academic Press, NY. 441 pp. Archer Daniels Midland Company. 1995. ADM Bioproducts Amino Acid Database. Archer Daniels Midland Company, Decatur, IL. Axe, D. E. 1994. Macrominerals. IMC-Agric Feed Ingredients, Bannockburn, IL. 50 pp. Bernhardt, M. D. 1996. What you need to know about feed grade minerals. Pp. 61–70 in 16th Annual Feed Ingredient Conference. Prince Agric. Products Inc., Quincy, Illinois. Centraal Veevoederbureau. 1994. Veevoedertabel. CVB Lelystad, The Netherlands. Cera, K. R., D. C. Mahan, and G. A. Reinhart. 1989. Apparent fat digestibilities and performance responses of postweaning swine fed diets supplemented with coconut oil, corn oil or tallow. J. Anim. Sci. 67:2040–2047. Cort, W. M., T. S. Vicente, E. H. Waysek, and B. D. Williams. 1983. Vitamin E content of feedstuffs determined by high-performance liquid chromatographic fluorescence. J. Agric. Food Chem: 31: 1330–1333. Cromwell, G. L. 1992. The biological availability of phosphorus in feedstuffs for pigs. Pig News Inform. 13: 75N–78N. Dale, N. 1995. Feedstuffs Ingredient Analysis Table—1995. Miller Publishing Co., Minnetonka, MN. Ewan, R. C. 1996. Energy Values of Feed Ingredients. 5th Revised Edition. Iowa State University, Ames, IA. 101 pp. Fickler, J., J. Fontaine, and W. Heimbeck. 1995. The Amino Acid Composition of Feedstuffs, Degussa Corporation, Ridgefield Park, NJ. Fonnesbeck, P. V., H. Lloyd, R. Obray, and S. Romesburg. 1984. IFI Tables of Feed Composition. International Feedstuffs Institute, Utah Agric. Exper. Stat., Utah State University, Logan, UT. 607 pp. Frigg, M. 1984. Available biotin content of various feed ingredients. Poult. Sci. 63: 750–753. Frigg, M., and L. Volker. 1994. Biotin inclusion helps optimize animal performance. Feedstuffs (January 3) 12–13. Heartland Lysine. 1995. Apparent Ileal Digestibility of Crude Protein and Essential Amino Acids in Feedstuffs for Swine-1995. Heartland Lysine, Chicago. IL. Institut National de la Recherche Agronomique. 1984. L’alimentation des animaux monogastriques: Porcs, Lapins, Volailles. Institut National de la Recherche Agronomique, Paris, France. Jondreville, C., J. Van den Broecke, F. Gatel, and S. Van Cauwenberghe. 1995. Ileal Digestibility of Amino Acids in Feedstuffs for Pigs. Eurolysine and ITCF Technical Institute for Cereals and Forages, Paris, France. 55 pp. Jongbloed, A. W. 1987. Phosphorus in the Feeding of Pigs: Effect of Diet on the Absorption and Retention of Phosphorus by Growing Pigs. Instituut voor Veevoedingsonderzoek, Lelystad, The Netherlands. 343 pp. Knabe, D. A. 1995. Survey of the content and digestibility of protein and amino acids in animal protein coproducts. Pp. 15–37 in North Carolina Swine Nutrition Conference. Raleigh: North Carolina State University. Ministry of Agriculture Fisheries and Food Standing Committee on Tables of Feed Composition. 1990. UK Tables of Nutritive Value and

Chemical Composition of Feedstuffs. Rowett Research Services LTD, Bucksburn, Aberdeen. 420 pp. National Research Council. 1982. United States–Canadian Tables of Feed Composition: Nutritional Data for United States and Canadian Feeds. 3rd ed. Washington, DC: National Academy Press. 772 pp. National Research Council. 1993. Nutrient Requirements of Poultry. 9th ed. Washington, DC: National Academy Press. 155 pp. Nelson, J. 1995. Trace mineral sources, quality, and biological availability. AFIA Feed Ingredient Institute, American Feed Industry Assoc., June 24–25th, Chicago, IL. Noblet, J., and Y. Henry. 1991. Energy evaluation systems for pig diets. Pp. 87–110 in E. S. Batterham, ed. Manipulating Pig Production III. Proc. 3rd Biennial Conference of the Australasian Pig Science Association, Albury, Australia. Noblet, J., H. Fortune, C. Dupire, and S. Dubois. 1993. Digestible, metabolizable and net energy values of 13 feedstuffs for growing pigs: effect of energy system. Anim. Feed Sci. Technol. 42: 131–149. Noblet, J., H. Fortune, X. S. Shi, and S. Dubois. 1994. Prediction of net energy value of feeds for growing pigs. J. Anim. Sci. 72: 344–354. North Central Region Committee on Swine Nutrition (NCR-42). 1992. Variability among sources and laboratories in chemical analysis of corn and soybean meal. J. Anim. Sci. 70(Suppl. 1):70 (Abstr.). North Central Region Committee on Swine Nutrition (NCR-42). 1993. Variability among sources and laboratories in selenium analysis of corn and soybean meal. J. Anim. Sci. 71(Suppl. 1):67 (Abstr.). North Central Region Committee on Swine Nutrition (NCR-42). 1995. Variability among sources and laboratories in analyses of corn and wheat middlings. J. Anim. Sci. 73(Suppl. 1):79 (Abstr.). Novus. 1994. Raw Material Compendium: A Compilation of Worldwide Data Sources. Novus International Inc., Brussels, Belgium. 541 pp. Pearl, G. G. 1995. Feeding Fats. Fats and Proteins Research Foundation Publication No. 269. Bloomington, IL: Fats and Proteins Research Foundation. 23 pp. Powles, J., J. Wiseman, D. J. A. Cole, and S. Jagger. 1995. Prediction of the apparent digestible energy value of fats given to pigs. Anim. Sci. 61: ˆ 149–154. Rhone-Poulenc. 1993a. Rhodimet Nutrition Guide: Feed ˆ Ingredients Formulation in Digestible Amino Acids, 2nd Edition. Rhone-Poulenc Animal Nutrition, Antony Cedex, France. 55 pp. ˆ ˆ Rhone-Poulenc. 1993b. Rhodimet Nutrition Guide, 6th Edition. RhonePoulenc Animal Nutrition, Antony Cedex, France. 39 pp. Roche. 1986. Biotin News No 5: Update on total and available biotin in feed ingredients. Roche Animal Nutrition and Health, Basel, Switzerland. 10 pp. Roche. 1987a. Beta-Carotene News: Beta-Carotene in Feedstuffs. Roche Animal Nutrition and Health, Basel, Switzerland. 18 pp. Roche. 1987b. Vitamin E News: Alpha-Tocopherol in Feedstuffs. Roche Animal Nutrition and Health, Basel, Switzerland. 42 pp. Roche. 1992. Folic Acid Content of Feedstuffs. Roche Animal Nutrition and Health, Basel, Switzerland. 9 pp. Southern, L. L. 1991. Digestible amino acids and digestible amino acid requirements for swine. BioKyowa Technical Review No. 2. 16 pp. U.S. Department of Agriculture, Agricultural Research Service. 1997. USDA Nutrient Database for Standard Reference, Release 11-1. Nutrient Data Laboratory Homepage, http://www.nal.usda.gov/fnic/ foodcomp.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 1 GROWTH MODEL

Equations Used to Model the Biological Basis for Predicting Nutrient Requirements

Calculations

Assumptions

1. Average FFL gain for the group of pigs is calculated from the number of barrows, gilts, and boars times the respective FFL gain rate divided by the total number of pigs in the group. 2. Carcass FFL gain, g/day. The fractional FFL gain based on the BW is calculated from the following relationship. Fractional carcass FFL gain 4 0.4767 ` (0.02147 2 BW) 1 (0.0002376 2 BW2) ` (0.000000713 2 BW3) The result is multiplied by the average FFL gain to give the carcass FFL gain for the simulation. 3. DE intake, Mcal/d 1 The predicted intake uses the following equations with adjustments for temperature, space per pig, and gender. Note: if feed intake is specified, the predicted feed intake is not used and the specified intake is used.

There are a number of constants or assumptions that are used in the model, and these are as follows. ● Rate of carcass fat free lean (FFL) gain is a function of body weight (BW) and is the same for barrows, gilts, and boars ● Lysine required per unit of protein accretion is 0.12 g of true digestible lysine/g of protein accretion ● Maintenance lysine requirement is 0.036 g/kg of BW raised to 0.75 power (BW0.75) ● ME is 0.96 2 DE ● Conversion of carcass FFL to whole body protein: whole body protein 4 FFL/2.55 ● Protein concentration in FFL tissue is 23% ● Fat concentration in fat tissue is 90% ● ME required for maintenance is 106 kcal/kg of BW0.75 ● ME required for protein synthesis is 10.6 kcal/g ● ME required for fat synthesis is 12.5 kcal/g

For pigs less than 20 kg BW: DE intake (kcal/day) 4 (251 2 BW) 1 (0.99 2 BW2 1 133 For pigs from 20 to 120 kg BW: DE intake (kcal/day) 4 1,250 ` (188 2 BW) 1 (1.4 2 BW2) ` (0.0044 2 BW3) 4. Adjustments to DE intake. a. Gender: between 30 and 120 kg BW, an adjustment for gender distribution is calculated from the following relationship. Adjustment (kcal/day) 4 DE intake 2 ((0.00385 2 BW) 1 (0.0000235 2 BW2) 1 0.083)

Input Data These include the dietary energy concentration (kcal of DE or ME/kg), daily DE or ME intake (Mcal; note that this value is optional, and if a value is entered it is used in subsequent calculations), pig weight (kg), the number of gilts and the potential FFL gain of gilts (g/day), the number of barrows and the potential FFL gain of barrows (g/day), the number of boars and the potential FFL gain of boars (g/day), space per pig in square meters (m2), and temperature in °C. The effects of space per pig are minimized at 0.41 m2 up to 20 kg BW, at 1.06 m2 from 20 to 50 kg BW, and at 1.10 m2 above 50 kg BW.

143

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

144

Nutrient Requirements of Swine

The adjustment is added for barrows and subtracted for gilts and boars. A weighting factor is calculated for mixed groups by subtracting the number of gilts and boars from the number of barrows and dividing by the total number of pigs. The adjustment is multiplied by the weighting factor to obtain the final adjustment to be added to DE intake. b. Space allowance (SP). The adjustment is calculated and the negative result is added to DE intake. 1. Pigs 20 kg BW or less: for space allowance less than 0.41 m2 DE intake adjustment 4 DE intake 2 (0.7227 ` (1.324 2 SP) 1 (1.5954 2 SP2) 11) 2. Pigs from 20 to 50 kg BW: for space allowance less than 1.059 m2 DE intake adjustment 4 DE intake 2 (0.7725 ` (0.4293 2 SP) 1 (0.2025 2 SP2) 11) 3. Pigs 50 kg BW or greater: for space allowance less than 1.095 m2 DE intake adjustment 4 DE intake 2 (0.6165 ` (0.7005 2 SP) 1 (0.32 2 SP2) 11) Note: Caution should be used in making adjustments for crowding because the adjustments are not precise. Adjustments may be too great at the lower end of each weight range category. c. Temperature (T). The adjustment is calculated and added to DE intake. Optimal Temperature (OT) 4 26 1 (0.0614 2 BW) % Change 4 (OT 1 T) 2 0.0165 Adjustment 4 DE intake 2 % Change 5. Potential whole body protein gain based on energy intake. Whole body protein gain based on the following equation. Whole body protein gain 4 2 1 2 2

(16.25 ` 17.5 e 10.0192 2 BW) (DE intake from step 3 (BW0.75 2 0.110 2 0.55) (1 ` (0.015 2 (20 1 T))) ((FFL gain/2.55)/125)

6. Carcass FFL gain (step 2) to whole body protein gain, g/day. Conversion of FFL gain to whole body protein gain by a factor of 2.55. Whole body protein gain 4 carcass FFL gain/2.55 7. Whole body protein gain (g/day) is the value that is used in the following calculations and is the smaller of the protein gain from step 5 and from the lean gain accretion curve (step 6).

8. True ileal digestible lysine required for maintenance (g/day). Lysine required for maintenance 4 0.036 2 BW0.75 9. True ileal digestible lysine required for protein gain (g/day). a. For pigs from 20 to 120 kg BW: True ileal digestible lysine for gain (g/day) 4 0.12 2 whole body protein gain (step 7) b. For pigs 20 kg BW or less: Total lysine, % 4 1.793 1 (0.0873 2 BW) ` (0.00429 2 BW2) 1 (0.000089 2 BW3) True ileal digestible lysine, % 4 (Total lysine 1 0.0365)/1.0973 Feed consumed, g 4 DE intake/DE concentration True ileal digestible lysine (g/day) 4 True ileal digestible lysine (%) 2 Feed consumed (g/day) 2 0.01 True ileal digestible lysine for gain (g/day) 4 True ileal digestible lysine (g/day) 1 Lysine required for maintenance (step 8) 10. True ileal digestible amino acid requirements (g/ day). The true ileal digestible lysine required for maintenance (step 8) is multiplied by the ratio of each amino acid to lysine for maintenance (Chapter 2, Table 2-1). The true ileal digestible lysine for gain (step 9) is multiplied by the ratio of each amino acid to lysine for gain (Chapter 2, Table 2-1). The sum of the requirement for maintenance and gain for each amino acid is the daily total true ileal digestible requirement. 11. True ileal digestible requirements (g/day) are converted to requirements as a percentage using the DE intake/day and the DE concentration of the diet. The true ileal digestible requirements as a percentage are converted to apparent or total using the equations in Chapter 3, Table 3-1. 12. Estimating potential performance a. ME intake ME intake 4 DE intake 2 0.96 b. ME required for protein synthesis Whole body protein gain (step 7) 2 10.6 c. ME for maintenance 106 2 BW0.75 d. Thermoregulatory energy Step 4c e. Energy available for fat synthesis ME intake 1 step b 1 step c 1 step d f. Fat synthesized (g/day) Step e/12.5 g. Fat tissue (g/day) Step f/0.9

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 1: Model Equations for Predicting Requirements h. Protein tissue (g/day) Whole body protein gain/0.23 i. Daily BW gain (g/day) (step g ` step h)/0.94 j. Feed-to-gain ratio Daily feed intake/Daily BW gain k. Crude protein (corn–soybean meal diet) Crude protein (%) 4 5.22 ` (15.51 2 True digestible lysine, %)

GESTATION MODEL Assumptions A number of constants or assumptions have been made and are as follows. ●

Gestation length is 115 days Gestation weight gain in the products of conception is 2.28 kg/pig (19.8 g/day for each pig) ● Protein gain in the products of conception is 245 g/ pig (2.13 g/day for each pig) ● Nitrogen gain in the products of conception is 39.2 g/pig (0.34 g/day for each pig) ● Body weight (BW) is the weight at breeding plus onehalf of the total gestation weight gain ● Metabolic body weight is BW0.75 ● ME required for maintenance is 106 kcal/kg of BW0.75 ● ME required for protein synthesis is 10.6 kcal/g ● ME required for fat synthesis is 12.5 kcal/g ● ME required for daily gain of the products of conception is 35.8 kcal/pig ● Lean tissue contains 23% protein ● Fat tissue contains 90% fat ● True ileal digestible lysine requirement for maintenance is 0.036 g/kg of BW0.75 ●

Input Data These include the DE or ME concentration (kcal/kg), daily DE or ME intake (optional) (kcal/day), sow weight at breeding (kg), the expected number of pigs born, the desired weight gain during gestation (kg) and temperature (°C). Note: Daily energy intake is an optional input. If daily energy intake is not provided, the energy required for the desired weight gain is calculated. If a value is entered for daily energy intake, weight gain is calculated using the input energy intake and then the amino acid requirements are calculated to support that level of performance.

145

Calculations 1. Maternal weight gain a. If DE or ME intake is not input, then: Maternal weight gain 4 Gestation weight gain 1 (2.28 2 No. of pigs) Maternal fat gain 4 (Maternal weight gain 2 0.638) 1 9.08 Maternal lean tissue gain 4 Maternal weight gain 1 Maternal fat gain b. If daily DE or ME intake is input then: ME in the products of conception 4 No. of pigs 2 35.8 kcal/pig Maintenance ME requirement 4 106 2 BW0.75 ME for maternal weight gain 4 ME intake 1 ME in the products of conception 1 maintenance ME Maternal weight gain 4 87 ` (ME for maternal weight gain 1 0.12171) Daily gestation weight gain 4 (Maternal weight gain) ` (No. of pigs 2 19.8 g) Gestation weight gain 4 Daily gestation weight gain 2 115 Because the maintenance requirement is dependent on gestation weight gain, these relationships are solved by iteration. Gestation weight gain is then partitioned into protein and fat gain by the same relationships as are used when daily DE intake is not provided. 2. Average daily nitrogen retention is the sum of maternal nitrogen retention and nitrogen in the products of conception. a. Maternal nitrogen retention N retention (g/day) 4 ((Maternal lean tissue)/115) 2 0.23 2 0.16 b. Nitrogen in the products of conception N retention (g/day) 4 No. of pigs 2 0.34 g/day 3. True ileal digestible lysine requirement for gain is calculated from N retention. True ileal digestible lysine (g/day) 4 (Total N retention 2 0.807) 4. True ileal digestible lysine requirement for maintenance. True ileal digestible lysine (g/day) 4 0.036 2 BW0.75 5. Total true ileal digestible lysine requirement is the sum of the requirement for gain and maintenance. 6. True ileal digestible amino acid requirement (g/day). The true ileal digestible lysine required for maintenance (step 4) is multiplied by the ratio of each amino acid to lysine for maintenance (see Chapter 2, Table 2-1). The true ileal digestible lysine required for gain (step 3) is multiplied by the ratio of each amino acid to lysine for protein accretion (Chapter 2, Table 2-1). The total require-

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

146

Nutrient Requirements of Swine

ment is the sum of the requirements for maintenance and gain. 7. True ileal digestible requirements (g/day) are converted to a percentage using the DE intake/day and the DE concentration of the diet. The true ileal digestible requirements as a percentage are converted to apparent or total using the equations in Chapter 3, Table 3-1. 8. Estimation of energy requirements a. ME for protein synthesis (Total N retained/0.16) 2 10.5 kcal/g b. ME for maternal fat synthesis Maternal fat gain 2 12.5 kcal/g c. ME for the products of conception No. of pigs 2 35.8 kcal/pig d. ME for thermoregulation. Because gestating sows are limit fed, only energy required to maintain body temperature is considered: (20 1 T) 2 BW 0.75 2 4.5. e. ME for maintenance 106 2 BW0.75 f. The total ME required is the sum of a through e. 9. Estimating potential performance a. Weight gain of the products of conception No. of pigs 2 19.8 g 2 115 days b. Maternal weight gain Gestation weight gain 1 (2.28 kg 2 No. of pigs) c. Maternal fat tissue gain ((Daily maternal weight gain 2 0.638) 1 9.08) 2 115 d. Maternal lean tissue gain Maternal weight gain 1 maternal fat gain e. Crude protein (corn–soybean meal diet) Crude protein (%) 4 5.22 ` (15.51 2 True digestible lysine, %)

Input Data These include the dietary DE or ME concentration (kcal/kg), daily DE or ME intake (optional) (mcal/day), sow weight after farrowing (kg), expected weight change during lactation (kg) (Note: weight loss is entered as a negative value), daily pig weight gain (g/day), lactation length (days), the number of pigs in the litter, and farrowing house temperature (°C). Note: Daily energy intake is an optional input value. If energy intake is entered, the weight change is calculated and any entered expected weight change is not used.

Calculations 1. is an a. b.

c. d. e. f. g. h. i. j.

LACTATION MODEL Assumptions A number of constants or assumptions have been made and are as follows: ● Body weight (BW) is the postfarrowing weight plus one-half of the total lactation weight change ● Metabolic body weight is BW0.75 ● ME required for maintenance is 106 kcal/kg of BW0.75 ● ME required for protein synthesis is 10.6 kcal/g ● ME required for fat synthesis is 12.5 kcal/g ● Lean tissue contains 23% protein ● Fat tissue contains 90% fat ● True ileal digestible lysine requirement for maintenance is 0.036 g/kg of BW0.75

Lactation weight change is calculated if DE intake input. ME intake 4 DE intake 2 0.96 (or input ME) ME required for milk 4 ((Daily weight gain/pig 2 No. pigs nursed 2 4.92) 1 (90 2 No. pigs nursed))/0.72 Maintenance ME requirement 4 106 2 BW0.75 Thermoregulatory ME requirement 4 (20 1 T) 2 310 Total ME required 4 Sum of steps b, c, and d ME available for maternal weight change 4 ME intake 1 Total required ME Protein weight change (g/day) 4 (((step f/0.88) 2 0.0942)/5.6) 1 1.47 Fat weight change (g/day) 4 ((step f/0.88) 1 ((step f/0.88) 2 0.0942))/9.4 Average daily weight change 4 (step g/0.23) ` (step h/0.9) Lactation weight change 4 Average daily weight change 2 Lactation length

Because the lactation weight change affects the metabolic body weight, and therefore the maintenance requirement, the series of equations are solved by iteration. 2. If DE or ME intake is not an input, the lactation weight change is calculated from the following steps: a. Average daily gain (ADG, g) 4 Lactation weight change (g)/Lactation length b. Protein in the weight change (g) 4 (ADG 2 0.0942) ` 1.47 c. Fat in the weight change (g) 4 ADG 1 (Protein in the weight change/0.23) d. Energy in the weight change 4 (Protein in the weight change 2 5.6) ` (Fat in the weight change 2 0.9 2 9.4)

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 1: Model Equations for Predicting Requirements

147

3. Lysine required a. Apparent digestible lysine for milk production (g) 4 (Daily pig weight gain 2 Pigs in litter 2 0.022) 1 6.39 b. Apparent digestible lysine for milk production (g/ day) is converted to a percentage using DE intake and DE concentration c. Percentage apparent digestible lysine is converted to percentage true digestible lysine True digestible lysine, % 4 1.050013 2 apparent digestible lysine, % ` 0.022052 d. Percentage true digestible lysine is converted to g/ day using DE intake and DE concentration e. Lysine for maintenance 4 0.036 2 BW0.75 f. Lysine from tissue 4 Change in protein 2 0.065 g. Total true ileal digestible lysine requirement 4 Sum of steps d, e, and f

DE concentration of the diet. The true ileal digestible requirements as a percentage are converted to apparent or total using the equations in Chapter 3, Table 3-1. 6. Estimation of energy partitioning a. ME for milk production 4 ((Daily pig weight gain 2 No. of pigs nursed 2 4.92) 1 (90 2 No. of pigs nursed))/0.72 b. Maintenance ME requirement 4 106 2 BW0.75 c. ME in body weight change 4 Daily ME change 2 0.88 d. Temperature adjustment: 20°C is considered the optimal temperature. Above the optimal temperature, ME intake is reduced 310 kcal/°C. Below the optimal temperature, ME intake is increased 310 kcal/°C e. Total ME required 4 Sum of steps a, b, c, and d f. DE intake 4 ME required/0.96

4. True digestible amino acid requirements (g/day). The true ileal digestible lysine required for maintenance (step 3e) is multiplied by the ratio of each amino acid to lysine for maintenance (see Chapter 2, Table 2-1). The true ileal digestible lysine required for milk production (step 3d) is multiplied by the ratio of each amino acid to lysine in milk protein (Chapter 2, Table 2-1). The lysine from tissue is multiplied by the ratio of each amino acid to lysine in tissue (Chapter 2, Table 2-1). The total requirement is the sum of the requirements for maintenance, milk production, and tissue change. 5. True ileal digestible requirements (g/day) are converted to a percentage using the DE intake/day and the

7. Estimating potential performance a. Maternal lean tissue change (Daily protein weight change (from step 1g or 2b)/ 0.23 2 Lactation length b. Maternal fat tissue change (Daily weight change 2 Lactation length) 1 Maternal lean tissue change c. Estimated milk production (kg/day) (ME for milk production [step 6a] 2 0.72)/1208 d. Crude protein (corn–soybean meal diet) Crude protein (%) 4 5.22 ` (15.51 2 True digestible lysine, %)

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 2 The lean growth rate of pigs is influenced by their total body growth rate and by the leanness of their carcass. Lean growth rate, or lean gain, is the rate at which a pig’s carcass accretes lean tissue (muscle). In the past, measurements of carcass lean tissue were standardized to 10 percent fat, then later they were standardized to 5 percent fat. Today, carcass lean tissue is measured on a fat-free basis. Carcass fat-free lean gain of pigs, in pounds per day, can be determined by estimating the pounds of fat-free lean in the carcass at the beginning and end of a prescribed test period, subtracting one weight from the other to determine the gain in weight of the fat-free lean, then dividing by the number of days from beginning to end. Final carcass fat-free lean, as a percent of the carcass (referred to as ‘‘fat-free lean index’’), can be estimated at the packing plant from either (1) ruler measurement of the backfat at the last rib at the midline of the hot carcass or (2) fat probe between the third and fourth rib, 7 cm off the midline of the hot carcass, according to the following (National Pork Producers Council, 1994):

Equations for Determining Lean Growth Rate of Pigs

A more sophisticated approach for measuring carcass fat-free lean weight and gain involves the determination of loin eye area in carcasses. If that measurement is obtainable, the lean gain can be calculated by the following (adapted from National Pork Producers Council, 1991): Carcass fat-free lean (lb) 4 0.95 2 [7.231 ` (0.437 2 hot carcass weight, lb) 1 (18.746 2 10th rib fat depth, in.) ` (3.877 2 10th rib loin eye area, in.2)] Initial carcass fat-free lean weight can be estimated by the following formula: Initial fat-free lean (lb) 4 0.95 2 [13.65 ` (0.418 2 live weight, lb)] After estimating the pounds of fat-free lean in the carcass at slaughter and in the carcass of the initial pig, the lean gain is calculated as follows: Carcass fat-free lean gain (lb/day)

Fat-free lean index 4 50.767 ` (0.035 2 Hot carcass weight, lb) 1 (8.979 2 Last rib midline backfat on hot carcass, in.)

4

(Final carcass fat-free lean, lb) 1 (Initial carcass fat-free lean, lb) Days from initial to final

Lean gain in lb/day is then converted to g/day by multiplying by 454.

or

Because lean gain is not constant from one day to the next, the calculated lean growth rate will be influenced by the initial and final weights of the pigs. The mean lean growth rate in the growth model assumes initial and final weights of 20 and 120 kg, respectively (approximately 45 and 265 lb). If the calculated lean growth rate is determined with pigs having different initial and/or final weights than these, the adjustment factors in Appendix Table 2-1 can be used to adjust the lean gain to a 20- to 120-kg basis.

Fat-free lean index 4 51.537 ` (0.035 2 Hot carcass weight, lb) 1 (12.260 2 Fat probe on hot carcass, in.) The pounds of carcass fat-free lean is then determined by multiplying the fat-free index (or percentage of fat-free lean in the hot carcass) by the weight of the hot carcass, as follows: Carcass fat-free lean (lb) 4 Fat-free lean index 2 Hot carcass weight (lb)

148

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 2: Determining Lean Growth Rate Appendix Table 2-1

Final Wt (kg) 40 45 50 55 60 65 70 75 80 85 90 95 100 105 110 115 120

149

Factors for use when the lean growth rate is measured over a period other than 20 to 120 kg.

20

25

30

35

40

45

50

55

1.07 1.05 1.03 1.02 1.01 1.00 0.99 0.98 0.98 0.98 0.98 0.98 0.98 0.98 0.99 0.99 1.00

1.04 1.02 1.01 1.00 0.99 0.98 0.97 0.97 0.97 0.97 0.97 0.97 0.97 0.98 0.98 0.99 0.99

1.02 1.00 0.99 0.98 0.97 0.96 0.96 0.96 0.96 0.96 0.96 0.96 0.96 0.97 0.97 0.98 0.99

1.00 0.99 0.97 0.96 0.96 0.95 0.95 0.95 0.95 0.95 0.95 0.95 0.96 0.96 0.97 0.98 0.98

0.97 0.96 0.95 0.95 0.94 0.94 0.94 0.94 0.94 0.94 0.95 0.95 0.96 0.97 0.97 0.98

0.95 0.94 0.94 0.94 0.93 0.93 0.94 0.94 0.94 0.95 0.95 0.96 0.97 0.97 0.98

0.94 0.93 0.93 0.93 0.93 0.93 0.94 0.94 0.95 0.95 0.96 0.97 0.98 0.99

0.93 0.93 0.93 0.93 0.93 0.94 0.94 0.95 0.95 0.96 0.97 0.98 0.99

Initial Wt (kg) 60 65

0.93 0.93 0.93 0.93 0.94 0.94 0.95 0.96 0.97 0.97 0.98 0.99

0.93 0.93 0.93 0.94 0.95 0.95 0.96 0.97 0.98 0.99 1.00

70

75

80

85

90

95

100

0.93 0.94 0.94 0.95 0.96 0.97 0.98 0.99 1.00 1.01

0.94 0.95 0.96 0.97 0.97 0.98 1.00 1.01 1.02

0.96 0.96 0.97 0.98 0.99 1.00 1.02 1.03

0.97 0.98 0.99 1.00 1.01 1.03 1.04

0.99 1.00 1.01 1.03 1.04 1.05

1.01 1.03 1.04 1.05 1.07

1.04 1.05 1.07 1.08

Instructions: Find the cell in the table that corresponds to the initial and final weights over which the mean lean growth rate in the situation of interest is measured. Multiply the measured mean lean growth rate by the factor in that cell. Note: These factors are correct only for the default lean accretion curve.

REFERENCES National Pork Producers Council. 1991. Procedures to Evaluate Market Hogs. Third ed. National Pork Producers Council, Des Moines, IA. National Pork Producers Council. 1994. Fat-Free Lean Index. National Pork Producers Council, Des Moines, IA.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Method to Create a Cubic Regression Equation

Appendix 3 This procedure is used to generate the coefficients for a cubic regression equation for a user-generated lean growth curve in the growth mode. These instructions apply to the Microsoft Excelt spreadsheet program. The use of this software program for this example should not be viewed as an expressed endorsement of the software by the authoring subcommittee or the National Research Council (see Note at the end of this Appendix). Other programs can be used, but the process of generating the coefficients will differ. The Y statistic in the regression equation is the percentage of the overall mean of the carcass fat-free lean growth rate (or mean protein accretion rate) at a given body weight, with this overall mean expressed as 1.00. For example, if a value is 80 percent of the overall mean, the value at that point is 0.80.

5. Carefully click on one of the data points; this should highlight all data points. If not, click in another area and try again. 6. From the menu at the top of the screen, click on ‘‘Insert,’’ and then select ‘‘Trendline.’’ 7. A Trendline dialog box will appear with ‘‘Type’’ and ‘‘Options’’ as two tabs. On the ‘‘Type’’ page, single click on the ‘‘Polynomial’’ box; then under ‘‘Order,’’ change to the number ‘‘3.’’ 8. Now click on the ‘‘Options’’ tab in the dialog box and click the small boxes beside ‘‘Display Equation on Chart’’ and ‘‘Display R-Squared Value on Chart.’’ Make sure that a check mark appears in those boxes. Click ‘‘OK.’’ 9. The equation is now shown on the graph (see Appendix Figure 3-1). If it is covered by the trend line, click on the equation and drag it to a clean area. 10. The coefficients shown in the equation are entered in the model under user-generated lean growth curve. Note that, for Excelt, the coefficients are in reverse order as compared with the coefficients in the model. The R2 value indicates how well the data points fit the trend line (an R2 value of 1.0 is a perfect fit). 11. If the fit is poor, you might try a quadratic trend line. Repeat the procedure, except change the ‘‘3’’ to a ‘‘2’’ in item 6. If you select a quadratic equation, rather than a cubic equation, then enter a ‘‘0’’ in the fourth blank in the user-generated lean growth equation in the growth model. 12. Print and Save.

1. Enter data so that the X variables (body weight) are in column A and the Y variables are in column B (or a column to the right of the X variables. NOTE: There must be at least 5 data points, and the mean of the Y variables should be 1.00. 2. Create a scatter plot of the data. First, highlight the data cells, then go to the tool bar and click on ChartWizard; the ChartWizard will guide you through the process if you follow the instructions in the dialog box at the bottom of the screen. 3. Click the mouse and drag a box on the screen. Then click in the following order: ‘‘Next, XY (scatter), Next, 1, Next, Next.’’ Under ‘‘Add a Legend?’’, click ‘‘No’’ then click ‘‘Finish.’’ 4. Double click on the chart that you have created; this will put you into chart edit mode.

Note: Excelt is a registered trademark of the Microsoft Corporation in the United States and/or other countries. The Microsoft Excel Solver program was developed by Frontline Systems, Inc., P.O. Box 4288, Incline

150

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 3: Creating a Cubic Regression Equation

151

Appendix Figure 3-1. Graph of a cubic regression equation. Village, NV 89450-4288. Portions of the Microsoft Excel Solver program code are copyright 1990, 1991, 1992, and 1995 by Frontline Systems, Inc. Portions are copyright 1989 by Optimal Methods, Inc. The Microsoft Excel Solver program uses Generalized Reduced Gradient (GRG2) nonlinear optimization code developed by Leon Lasdon, University of Texas at Austin, and Allan Waren, Cleveland State University. Linear and integer problems use the simplex method with bounds on the variables and the

branch bound method, implemented by John Watson and Dan Fylstra, Frontline Systems, Inc. The Microsoft Excel Analysis Toolpak was developed by GreyMatter International, Inc., 173 Otis Street, P.O. Box 388, Cambridge, MA 02141. The Microsoft Excelt Spreadsheet Solution Templates were developed by Village Software, 186 Lincoln Street, Boston, MA 02111.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 4

NUTRIENT REQUIREMENTS OF SWINE Tenth Revised Edition, 1998

A USER’S GUIDE FOR MODEL APPLICATION

Subcommittee on Swine Nutrition Committee on Animal Nutrition Board on Agriculture National Research Council

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

SUBCOMMITTEE ON SWINE NUTRITION GARY L. CROMWELL, Chair, University of Kentucky DAVID H. BAKER, University of Illinois RICHARD C. EWAN, Iowa State University E. T. KORNEGAY, Virginia Polytechnic Institute and State University AUSTIN J. LEWIS, University of Nebraska JAMES E. PETTIGREW, Pettigrew Consulting International, Louisiana, Missouri NORMAN C. STEELE, U.S. Department of Agriculture, Agricultural Research Service, Beltsville, Maryland PHILIP A. THACKER, University of Saskatchewan, Canada STAFF CHARLOTTE KIRK BAER, Program Director MELINDA SIMONS, Project Assistant SOFTWARE INTERFACE DEVELOPMENT RONALD HAUGEN, Easy Systems, Inc., Welcome, Minnesota

154

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Table of Contents

Chapter 1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156 Tutorials, 156 Computer Model Programs, 156 Feed Composition Tables, 157 Hardware and Software Requirements and Program Installation, 157 Getting Started, 158 Chapter 2. Using the NRC Model Programs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159 About the Screen, 160 Task Field, 160 Report Field, 161 Folder Field, 161 Chapter 3. Tutorial Lesson 1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 General Description of the Growth Model, 163 Grow-Finish Case Study, 164 Chapter 4. Tutorial Lesson 2. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166 General Description of the Gestation Model, 166 Gestating Sow Case Study, 166 Chapter 5. Tutorial Lesson 3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 General Description of the Lactation Model, 167 Lactating Sow Case Study, 167

155

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

1 A compact disk (CD-ROM) containing three models is provided as a companion to the National Research Council’s (NRC’s) Nutrient Requirements of Swine, Tenth Revised Edition, 1998. This User’s Guide provides information with specific examples demonstrating the use of each of the three programs contained on the NRC computer disk. Three model program options for predicting nutrient requirements include programs that address (1) the growing pig, including the starting, growing, and finishing stages of growth from 3 to 120 kg body weight, (2) gestating sows, and (3) lactating sows. These models allow the user to apply the modeling principles, interrelationships, and equations summarized in Chapter 3 of the report. A basic understanding of swine nutrition is required to properly use the programs, and knowledge of the underlying biological concepts presented in this report is essential for appropriate use of the computer models. The computer model programs predict energy and amino acid requirements on a daily basis and on a diet concentration basis by estimating the amount of energy and amino acids needed for specific physiologic functions, such as maintenance, body protein accretion, and milk protein synthesis. Equations are given to predict mineral and vitamin requirements at various stages of growth. Numerous factors including body weight, lean growth potential, gender, environmental temperature, space per pig, number of pigs per litter born and weaned, weaning age, litter growth rate, and other factors enter into the equations for predicting nutrient requirements. Each program uses different prediction equations and the inputs required for each program vary according to the various physiologic and metabolic functions being described (i.e., growth, gestation, lactation). The software was developed for accuracy and ease of use. The programs were developed in database format as the foundation for a Windows-based, menu-driven program. Program ‘‘help’’ screens provide guidelines for choosing inputs and in interpreting and applying outputs.

Introduction

TUTORIALS The purpose of this user’s guide is to demonstrate how to apply the NRC computer model programs to predict nutrient requirements of swine. Tutorials familiarize the user with program mechanics and options. In addition, they provide a quick overview of the program applications for each of the three models. Examples are provided that allow the user to input data, obtain predicted nutrient requirements, and evaluate the results. The user is strongly urged to read the comprehensive material provided in this report and is referred to the following chapters for detailed information on biological bases for equations and assumptions used in the software: ● ● ●

Energy, Chapter 1 Proteins and Amino Acids, Chapter 2 Models for Estimating Energy and Amino Acid Requirements, Chapter 3 ● Minerals, Chapter 4 ● Vitamins, Chapter 5 ● Diet Formulation, Chapter 9 ● Nutrient Requirement Tables, Chapter 10 ● Composition of Feed Ingredients, Chapter 11

COMPUTER MODEL PROGRAMS Growth This model allows the user to compute the daily lysine requirement on a true ileal digestible basis for maintenance and whole-body protein accretion. The program uses a lean tissue accretion curve, based on carcass fat-free lean tissue, to predict the lean tissue accretion rate at a given body weight. The carcass lean tissue accretion is then converted to whole body protein accretion by assuming there are 2.55 grams of carcass fat-free lean tissue per gram of whole-body protein. The amount of true ileal digestible

156

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 4: User’s Guide lysine needed to support that amount of whole body protein accretion is then calculated, based on the assumption that 0.12 gram of lysine is needed for every gram of whole body protein. This amount of lysine needed for protein synthesis plus the amount needed for maintenance are summed. The requirements for the other amino acids are based on the ideal protein system; that is, their ratio to lysine for maintenance and protein accretion. Equations are used to estimate digestible energy intake based on body weight, gender, environmental temperature, and space per pig. Feed intake is then determined from the estimate of digestible energy intake. Mineral and vitamin requirements are estimated for different body weights by equations. Gestation This model predicts the amount of dietary energy and amino acids needed by sows of different breeding weights to attain a targeted weight gain during pregnancy. The formula for prediction assumes that total tissue accretion is the sum of that in the maternal body plus the products of conception. The model predicts energy requirement for a given gestation weight gain, or predicts weight gain resulting from a given energy intake. Adjustments in energy intake are made for cold environments. Lactation This model predicts the amount of dietary energy and amino acids needed by sows based on their postfarrowing weight, the weight loss or gain during lactation, and the weight gain of the litter, a reflection of the sow’s level of milk production. Amino acid requirements are based on the amino acid patterns in tissue protein gain or loss, milk protein, and maintenance. The model predicts energy requirements based on the sow’s lactational weight change, or predicts lactation weight change resulting from a given energy intake.

FEED COMPOSITION TABLES These tables are provided to allow the user to view and print the composition of 79 different feed ingredients that are commonly used in swine feeding. The tables are also presented in Chapter 11. Adobe Acrobat Reader is required to view and print the tables. Instructions for installing Adobe Acrobat Reader are provided in the section ‘‘Getting Started’’ in this chapter (Chapter 1) of the User’s Guide. After Adobe Acrobat Reader has been installed, the tables can be viewed by pressing ‘‘Tables’’ at the top of the screen (or if you are already in one of the model programs, by pressing ‘‘Feed Tables’’ at the bottom of the

157

screen). The first screen to appear provides titles of each of the 11 tables. To access a particular table, click on its title and the table containing feed composition data will appear. To move around within individual tables, use the cursor (designated on the screen as a small hand, which points to certain locations as the mouse is moved to different areas on the table). Various tool tips located on the tool bar at the top of the screen allow the user to move from page to page, to display entire tables on the screen, and to conduct other procedures. For instance, to zoom in on a certain area of the table, select the tool designated as a magnifying glass with a ‘‘`’’ in its center. Move the magnifying glass to the desired section of the screen and click. The portion of the screen being viewed will become magnified. By continuing to click, the image will continue to be enlarged. Portions of the table can also be magnified by selecting with the mouse while holding the mouse button down and outlining the area desired. When the mouse button is released, the area becomes enlarged. To return to the full view, click on the button that shows a full page on the tool bar. Three buttons located on the left side of the tool bar provide the following useful options: the first button displays only the page, the second button from the left provides a list of tables in the margin of the screen, and the third button from the left provides a full-page view of several tables in the margin. To exit the tables and return to the main menu, select ‘‘File’’ at the top of the screen and then ‘‘Exit.’’ An alternate way to exit is to hold down the ‘‘Alternate’’ key on the keyboard and press F4. Risk of use: Because of the many variables involved and judgments that must be made in choosing inputs and interpreting outputs, the NRC makes no claim for the accuracy of this software and the user is solely responsible for risk of use.

HARDWARE AND SOFTWARE REQUIREMENTS AND PROGRAM INSTALLATION This software is designed to operate in a Windows environment on microcomputers that run Windows 3.1 or higher versions (Windows 95 or NT). The NRC model requires the following hardware: 1. an IBM compatible computer with 80386sx processor (or higher) with mouse; 2. 8 Megabytes random access memory (RAM) (16 Megabytes recommended); 3. 16 Megabytes hard drive disk space; and

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

158

Nutrient Requirements of Swine

4. a compact disk (CD-ROM) drive (internal or external) or floppy-disk drive. The NRC model requires the following software: 1. Windows version 3.1 or higher, Windows for Workgroups 3.11 (Windows 95 or NT recommended), 2. NRC model CD or floppy disks.

To install this software from floppy disks: 1. 2. 3. 4. 5. 6. 7. 8.

Enter Windows. Select Program Manager. Select File. Select Run. Insert NRC ‘‘Disk 1’’ into floppy drive. Type ‘‘A: ⁄ SETUP’’ on the command line. Press Enter. Insert disks as prompted.

To install this software from the CD: 1. Insert CD into drive. Setup will begin automatically if you are using Windows 95. If not, locate the ‘‘setup’’ folder (directory) named ‘‘Disk 1.’’ 2. If you are using Windows 3.1, you will need to locate the ‘‘setup’’ folder (directory) named ‘‘Disk 1’’ on the CD to initiate the setup.

To copy the program from the CD to floppy disks: 1. General instructions: Insert the CD. Locate the 13 disk folders (directories) on the CD. Copy all files in each folder to individual disks inserted into your floppy drive. You will need 13 individual floppy disks. You may also copy the contents of the CD to a directory on your hard drive and subsequently copy the 13 folders onto individual disks inserted in your floppy drive. 2. Using Windows 95: Use Explorer to locate your CD drive. Click on folder named ‘‘Disk 1.’’ On the right-hand side you will see all files contained in this folder. Press ‘‘Alt-A’’ to select all files in this folder. Next, make sure your mouse is located over the selected files and RIGHT CLICK your mouse. Select ‘‘Copy’’ and move to the disk drive A: and PASTE onto the floppy disk located in drive A. Label this as ‘‘Disk #1.’’ Repeat for each of the 13 disks, labeling each with the appropriate disk number. 3. Using Windows 3.1: Use File Manager to copy as described above.

GETTING STARTED 1. After you have installed the NRC model, an icon will be added to your workspace, if you are using Windows 3.1. If you are using Windows 3.1, double click on that icon to open the model software. If you are using Windows 95, select ‘‘Programs’’ and then select ‘‘NRC’’ to begin. 2. An introductory screen will appear. After reading the statement of use, click ‘‘OK’’ 3. When the main menu screen appears, choose one of the program options on the screen (Gestation, Lactation, Growth) to select the desired program. Note: preferences for the use of either digestible energy (DE) or metabolizable energy (ME) for inputs should be selected as well as preference for the use of ‘‘tool tips’’ (use of these tips is recommended). Preferences on this main menu screen should be selected before beginning the model programs. 4. To view and print the feed composition tables, Adobe Acrobat Reader is required. To activate this software from the CD (or if the software currently exists on your computer) click on the utilities button located on the menu bar of the NRC model. You will be presented with the option of installing or providing the location of Adobe Acrobat Reader. Select the installation option to activate Acrobat Reader. Once this software is installed, click on the ‘‘tables’’ button on the main menu bar to view and print the feed composition tables. You can also access and download Adobe Acrobat Reader software onto your computer by going to the Internet address: www.adobe.com

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

2 Program features of the NRC model are presented in the following examples. These tutorials describe how to

Using the NRC Model Programs

choose inputs and how to obtain, interpret, and apply outputs.

NRC Model Program MAIN MENU SCREEN

159

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

160

Nutrient Requirements of Swine

Position the cursor over the appropriate program option and click or press ^ENTER& to select that program.

ABOUT THE SCREEN

can be viewed as having three functional sections. Along the bottom is the task field with a series of task bars (Figure 1). Above the task field you will find the screen is divided into two equal sections; the report field on the right (Figure 2), and the folder field on the left (Figure 3).

From the main menu select the NRC model Growth program. The user will be presented with a screen that

TASK FIELD

FIGURE 1 Screen for the Growth program with task field highlighted.

Reports Allows the user to view the results of the program computations. The user can view input data, calculations, and nutrient requirements specific to the program used. Reports are given for basic calculations, true ileal digestible amino acids, apparent ileal digestible amino acids, total amino acids in a corn–soybean meal diet, minerals, vitamins, and fatty acids, by clicking on the A, B, C, D, E, F, or G button at the bottom of the report field, respectively. The report heading button allows the user to input a title for the report printout.

Min/Max Changes the screen in the report field to allow input of new minimum and maximum parameter values. Within this screen, the user can also change the incremental step by which increases and decreases in values can be made using the arrow functions or the keyboard arrows. The user can also set the number of decimal places appearing in a parameter. After a change is made in the input range, incremental step, or decimal places, the appropriate save button becomes activated. Changes must be saved before moving to another parameter in order to retain the change.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 4: User’s Guide Conversions Changes the screen in the report field to display the U.S. standard equivalents to the units displayed in the folder field. Enter Dietary Energy Intake Allows the user to calculate the nutrient requirements based on an actual energy intake or on the model’s default energy intake. Report Based On Allows the user to view the nutrient requirements on an amount-per-day basis or on a diet-concentration basis. Auto Calc—Off Allows the user to determine whether to have the program recalculate requirements after each input, or only after manually activating recalculation function. (Recommend setting to the ‘‘off’’ position when using slower machines.) Reset Returns all parameters to their original default values. Feed Tables Allows the user to access the feed composition tables. Help Allows the user to access help screens. Preview Allows the user to preview the report results of the model being used. This option provides a view of results that occur from changing inputs. The report includes information provided by the user to generate the report and the calculated requirements on both a percentage basis and an amount-per-day basis. Print Sends the current version of the report associated with the model program being used to the printer selected in the Windows control panel.

161

Exit Allows the user to exit the computer model and return to the main menu.

REPORT FIELD This field contains three separate functional screens. The user may toggle between the three screens by clicking on either the Report, Min/Max, or Conversions buttons. Report Screen The report screen allows the user to choose the view of one of seven different reports by clicking on the appropriate button. A Basic calculations and input data B

Amino acids on a true ileal digestible basis

C

Amino acids on an apparent ileal digestible basis

D

Amino acids on a total basis (applies to corn–soybean meal diet)

E

Minerals

F

Vitamins

G

Fatty acids

Report Heading Report Heading allows the user to define the title of the report. Parameter Setting Screen The parameter setting screen allows the user to change values for parameter inputs. This screen is accessed by clicking the Min/Max button. Conversions Screen Provides the user with standard U.S. equivalents to metric units shown in the folder field. This screen is accessed by clicking the Conversions button.

FOLDER FIELD Within the folder field, the user can move between the gestation, lactation, and growth folders by clicking on the appropriate folder tab at the top.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

162

Nutrient Requirements of Swine

FIGURE 2 Screen for the Growth program with report field highlighted.

FIGURE 3 Screen for the Growth program with folder field highlighted.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

3 GENERAL DESCRIPTION OF THE GROWTH MODEL

Tutorial Lesson 1

Optional Inputs The first entry is to indicate the caloric density of the diet, in kcal of DE per kg of diet assuming that DE was selected as the preferred input on the opening screen. Insert this value in the upper box. The default value is 3400 kcal/kg, which is the typical value for a corn–soybean meal diet. The user should then indicate the weight of the pig. The model was developed for pigs weighing from 3 to 120 kg, and any weights outside this range may not have valid estimates of requirements. Options exist for determining requirements of gilts, barrows, boars, or any combination of the three genders. This is attained by inserting the number of each sex in the appropriate box beside gilts, barrows, and boars. The model uses a standard default lean growth curve for gilts, barrows, and boars, and this is done by pressing the radio button beside ‘‘std.’’ Ordinarily, this is the lean growth curve that will be used. However, if the user prefers to use a different shape curve, that can be done also by pressing the button beside user. A new screen appears where you can enter the intercept and the coefficients of a cubic equation that best describes your alternative curve. A quadratic equation can also be used by using a zero as the fourth coefficient. Quadratic or cubic equations can be derived by using a spreadsheet program (See Appendix 3). Any new equations can be saved by pressing the appropriate buttons on this screen. Press ‘‘Return’’ to return to the main screen. To incorporate the effect of stocking density (space per pig) or environmental temperature, click on the appropriate box. When successfully incorporated, a check mark will appear in the box. To remove the effects of space per pig or environmental temperature, click again on the box to remove the check mark. The estimated requirements can be calculated with the option of either including the dietary energy intake or by excluding it and allowing the model to determine a default

All programs, including the Growth program, can be initiated by selecting the appropriate NRC model button on the opening screen. However, before selecting the appropriate model button, DE or ME should be selected as your preference of input values. Once in the desired model program, using your mouse or tab key, you can begin entering inputs (Figure 1). To move from one input cell to another, press the ‘‘Enter’’ or ‘‘Tab’’ keys.

Inputs To change values for inputs, highlight the appropriate cell by sweeping over the cell with the cursor or by tripleclicking on the cell, input the desired value, and hit enter. When the cursor is over the cell, a minimum and maximum value for input will appear. Values outside this range will not be accepted. Small incremental changes may be accomplished by one of two methods. After highlighting the cell incremental increases or decreases in the value can be made using the up or down arrows on the keyboard. Alternatively, these changes can be made by clicking on the up or down arrows to the right of the value in the appropriate cell. After you have attained the desired value, hit enter to move the cursor to the next input cell. Three other methods exist for moving between input cells. Any cell may be reached directly using the mouse cursor. You can move from one cell to the next using the tab key to move forwards, or use shift`tab to move backwards. Alternatively, the same task can be achieved using the left arrow key on the keyboard to move backwards and the right arrow key to move forwards. It is very important to remember to hit the Re-Calc task bar every time that an input is made. Otherwise, the report will not be correct. If the Re-Calc bar is a red color, this means that an input change has occurred and that the equations must be recalculated.

163

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

164

Nutrient Requirements of Swine

energy intake at a particular body weight. To change this from the default, click on the ‘‘Yes’’ option. Similarly, the program offers the option of having the reports given on a dietary-concentration basis or on an amount-per-day basis. The default setting is for a dietary-concentration basis. To obtain requirements on a daily basis, click on amount/day. At any time the user can return to the default settings by clicking on the ‘‘Reset’’ task bar. Options exist allowing the user to view the general output or the requirements for amino acids on a true or apparent digestible basis or on a total basis, minerals, vitamins, and fatty acids by clicking on boxes A through G at the bottom of the report screen. Conversion Factors In addition to these features, the standard English equivalent conversions of the metric values can be viewed by clicking on the ‘‘Conversions’’ task bar in the task field. To return to the report screen click on the ‘‘Report’’ bar. Maximum and Minimum Values To view the minimum and maximum values that can be input, click on the ‘‘Min/Max’’ bar in the task field. To view the range for any parameter, click on the parameter cell of interest. The range of acceptable values can be changed. To change the value for any given parameter, click on either the minimum or maximum bar in the box labeled ‘‘input ranges’’ and input the new value. Save the new values before exiting the program. Within this screen, there also exists the option to increase or decrease the degree of incremental changes when using the arrows button option. The number of figures following the decimal point for any parameter can also be changed within this screen.

GROW-FINISH CASE STUDY Begin the tutorial by selecting DE input values and then selecting the NRC Growth program. Both of these options are located on the main menu screen. At the Growth model working screen, enter information by placing the cursor in the appropriate cell. First, if the Re-Calc task bar says that it is on, turn it off. Otherwise, the program will recalculate between each entry and it will take several seconds for each recalculation. When you click on each input cell, the descriptor becomes highlighted; type information in highlighted cells. It may be necessary to sweep across the value with your mouse before entering the new number; or you can use the up and down arrows for minor changes. All entries are in metric units; conversions and conversion factors are available by clicking on the Conversion option,

which displays information and converted equivalents in English units. Dietary energy concentration Enter the dietary digestible energy (DE) concentration of the diet in kcal of DE per kg of diet. This is on an ‘‘as fed’’ basis and assumes 90% dry matter in the diet. In this example, the dietary energy concentration is 3,400 kcal of DE/kg. For practical results, this value cannot be below 2,000 or above 5,000 kcal/kg. The default value for this entry is 3,400. Daily energy intake In this example, we will not enter DE intake, but let the model calculate it from the default equation. Press the radio button beside ‘‘no’’ in the lower left task bar. Pig weight Enter the average weight of the pigs for which requirements will be determined. For this example, the average weight is 70 kg, so enter 70. Distribution of sex In this example, we will determine the requirement for a mix of barrows and gilt, assuming a 1:1 ratio. Enter a ‘‘1’’ in the space following gilts, a ‘‘1’’ in the space following barrows, and a ‘‘0’’ in the space following boars. If you enter 10 gilts and 10 barrows, you will get the same answer. Lean gain The average lean growth rate of the pigs is now entered. Lean gain is defined as the carcass fat-free lean tissue gain averaged over the range of 20 to 120 kg body weight. This value is estimated from the final lean content of the carcass minus an initial lean content of the carcass divided by the number of days from start to finish. In this example, we will assume that the lean gain is 330 grams per day for gilts and 320 grams for day for barrows. Enter these two values beside gilts and barrows, respectively. It does not matter what value is entered beside boars, because we are not including any boars in this determination (i.e., the number of boars is 0). Lean growth curve We will use the default growth curve for both genders, so press the radio button beside Std for barrows and for gilts. Space/Pig We will enter the space allocation, so click on the box to the left of space/pig. Let’s assume that there is 0.9 square meters per pig. Enter 0.9 in the box following space/pig. Note: Caution should be used in making adjustments for crowding because the adjustments are not precise. Adjustments may be too great at the lower end of each weight range category. Temperature We will enter 22 degrees C as an average temperature for these pigs. Check the box to the left of

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 4: User’s Guide temperature and enter 22 in the box to the right of temperature. Click on the box marked Conversions, and you will notice that the space in more familiar units is 9.68 square feet and the temperature is 71 degrees Fahrenheit. Click on the box called ‘‘Reports’’ to return the main screen. Go to the box marked ‘‘Report based on’’ and click the radio button next to ‘‘%’’ or ‘‘amount/kg.’’ You are now ready to look at the report. But, remember, it will not be correct until you first hit the Re-Calc task bar, which is red in color. Press it and after a few seconds, the task bar will say ‘‘Auto Calc Off.’’ Press box ‘‘A’’ under the report. You will now see much of the data that you have already entered. In addition, you will see the carcass daily lean growth rate and the whole body daily protein accretion rate of the pigs at the particular weight that you have chosen. The estimated whole body gain of protein tissue and fat tissue is also displayed. You will also see the DE and ME concentrations of the diet, the daily DE and ME intakes under the conditions that you have specified, the daily feed intake, and the expected daily gain and feed conversion efficiency (feed/gain) of the pigs at this particular body weight. The approximate crude protein level in the diet is also given. You will have to use the scroll bar on the right side of the screen to see the crude protein level of the diet. Click on ‘‘B’’ and the amino acid requirements, on a true ileal digestible basis, are listed. Click on C for the apparent ileal digestible amino acid requirements, and click on D for the total amino acid requirements assuming that a cornsoybean meal diet is fed. Note that the total lysine requirement is 0.75% of the diet. Click on E, F, and G for the mineral, vitamin, and linoleic acid requirements, respectively. Now click once again on D. Again, what you see are the amino acid requirements of a group of barrows and gilts (1:1 ratio) at 70 kg body weight with lean growth rates as you previously specified. Now go to the box marked ‘‘Report based on’’ and click on the radio button to the left of ‘‘amount/day.’’ Note that the report screen now shows that the daily lysine requirement is 19.7 grams/day. Now go back and click on % or amount/kg.

165

Let’s make some further changes to see what effect they have on the lysine requirement. Change the temperature, space/pig, lean growth rate of the pigs, number of pigs and you will see the resulting changes in the predicted lysine requirement. Click on the ‘‘Yes’’ button in the box marked ‘‘Enter Dietary Energy Intake’’ and reenter another daily energy intake. Note the change in the lysine requirement, as well as all of the other requirements when they are expressed on a concentration basis. This is because the feed intake changes; thus the concentration of nutrients must change to give an equivalent daily amount of that nutrient. Note that when you make a change in energy concentration, space per pig, or temperature, a warning sign will appear. The warning indicates that with low energy diets, pigs may not be able to eat sufficient feed to meet their predicted energy requirement and that with crowding or high temperatures, lean gain may be reduced. Note that every time you make a change in any of the variables, the Re-Calc task bar below the report will be in bright red, and you must click on this bar in order for the model to make recalculations. The recalculation box can be turned on or off. When turned on, recalculation occurs every time you make a change in your input, but it takes more time that way. When turned off, time-consuming recalculations are avoided whenever you change an input, but you must remember to click on the box to recalculate. Press the ‘‘Print’’ task bar and the report will be printed. If you want to assign a title to the report before printing it, press the ‘‘Report Heading’’ button and entire the desired title. The printed report for the conditions that you have entered is shown at the end of this exercise. If you want to view the report before it is printed, press the preview bar. Anytime that you need help, hit the ‘‘Help’’ task bar and a help screen will appear. The index will first appear and you can click on the area that you need help and that area will appear. Or, you can scroll through the entire help screen. If you want to exit the program, press ‘‘exit.’’ If you want to access one of the other models, click on the appropriate tab at the top left of the screen.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

4 GENERAL DESCRIPTION OF THE GESTATION MODEL

Tutorial Lesson 2

Assume that the sow will farrow 13 pigs, so make that entry. In this instance, we will assume that there is no temperature effect, so the small check box beside temperature should be blank. If there is already a value in the temperature box, click on the check mark beside it and the temperature box will go blank. Press the red ‘‘Re-Calc’’ button to make the calculations. Press the ‘‘A’’ box, and you will see the data that you entered. You will also see estimates of how the weight gain will be distributed, that is, 29.64 kg for the pigs and the placental tissues and 15.36 kg for maternal tissues. Of the maternal tissue, 14.64 kg is protein or lean tissue, and only 0.72 kg is fat tissue. The model predicts that the sow will consume 8809 kcal of DE per day, which is equivalent to 2.13 kg of feed per day. Click on boxes B, C, and D for estimates of the amino acid requirements of this sow, on a true digestible basis, apparent digestible basis, and total basis, respectively. Note that the model predicts that the sow requires 0.52% total lysine in the diet. Press boxes E, F, and G to see the mineral, vitamin, and linoleic acid requirements. Go to the ‘‘Report Based On’’ task bar and press ‘‘Amount/ day.’’ Now look at the reports after pressing B, C, D, etc. and you will see estimated requirements expressed on a daily basis. What would happen if this sow were fed either more or less feed during gestation? Go to the task bar marked ‘‘Daily Dietary Energy Intake’’ and press ‘‘Yes.’’ You can now enter DE intake in Mcal. Remember that there are 1000 kcal in 1 Mcal. Enter a different value, press the ‘‘ReCalc’’ bar and observe the changes in pregnancy weight gain and nutrient requirements that result from this change. To print the report, press ‘‘Print.’’ If you want to assign a title to the report before printing it, press the ‘‘Report Heading’’ button and entire the desired title. An example of a printed report based on the original entries is shown in the attached gestation table.

The Gestation model is quite similar to the Growth model in terms of data entry and output of reports. The user is referred to the general description of the Growth model, discussed in the previous chapter for details.

GESTATING SOW CASE STUDY Begin the tutorial by selecting the Gestation tab if you are already in one of the other programs or select Gestation from the main menu screen. Enter information by placing the cursor in the appropriate cell. First, if the Auto Calc is on, click on the task bar to turn it off. Otherwise, the program will recalculate between each entry and it will take several seconds for each recalculation. When you click on each input cell, the descriptor becomes highlighted; type information in highlighted cells. It may be necessary to sweep across the value with your mouse or triple-click on the cell before entering the new number; or you can use the up and down arrows for minor changes. All entries are in metric units; conversions and conversion factors are available by clicking on the Conversion option, which displays information and converted equivalents in English units. Enter the energy concentration of your diet. Let’s assume that you are feeding a diet with 3,200 kcal of DE per kg, which is slightly lower in energy than a corn–soy diet. Enter 3200. Daily energy intake will be calculated by the model. If there is a value in this box, then go the lower part of the screen and find Enter Dietary Energy Intake and click on the ‘‘no’’ box. The energy intake entry box will now be blank. Let’s assume that the sow weighs approximately 180 kg at breeding and we want it to gain 45 kg of weight during pregnancy. Enter 180 and 45 in the appropriate boxes.

166

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

5 GENERAL DESCRIPTION OF THE LACTATION MODEL

Tutorial Lesson 3

will be nursing 10 pigs, and the average daily gain per pig will be 200 grams. Enter these values in the appropriate boxes. Let’s assume that there will be no temperature effect so the small box beside temperature should not be checked. If there is already a value in the temperature box, click on the check mark beside it and the temperature box will go blank. Press the red ‘‘Re-Calc’’ button to make the calculations. Press the ‘‘A’’ box and you will see the data that you entered. You will also see that the sow will probably lose about 1.9 kg of lean tissue and 3.1 kg of fat tissue during the 21-day lactation, and it will be expected to produce about 7.4 kg of milk per day. The model predicts that the sow will consume about 16,942 kcal of DE per day, which is equal to 4.98 kg of feed. Click on boxes B, C, and D for estimates of the amino acid requirements of this sow, on an true digestible basis, apparent digestible basis, and total basis, respectively. Note that the model predicts that the sow requires 0.94% total lysine in the diet. Press boxes E, F, and G to see the mineral, vitamin, and linoleic acid requirements. Go to the ‘‘Report Based On’’ task bar and press ‘‘Amount/ day.’’ Now look at the reports after pressing B, C, D, etc. and you will see estimated requirements expressed on a daily basis. What would happen if this sow were fed more feed during lactation? Go to the task bar marked ‘‘Daily Dietary Energy Intake’’ and press ‘‘Yes.’’ You can now enter DE intake in Mcal. Remember that there are 1000 kcal in 1 Mcal. Enter 18.5 Mcal, press the ‘‘Re-Calc’’ bar, and notice the increased feed intake. The sow now will gain about 1.1 kg of weight during lactation and the percentages of amino acids are less (the lysine requirement is now 0.90%) because feed intake is higher. To print the report, press ‘‘Print.’’ If you want to assign a title to the report before printing it, press the ‘‘Report Heading’’ button and entire the desired title. An example of a printed report based on the original entries is shown in the following lactation table.

The Lactation model is quite similar to the Growth model in terms of data entry and output of reports. The user is referred to the general description of the Growth model, discussed in Chapter 3 of this User’s Guide for details.

LACTATING SOW CASE STUDY Begin the tutorial by selecting the ‘‘Lactation’’ tab if you are already in one of the other programs or select ‘‘Lactation’’ from the main menu screen. Enter information by placing the cursor in the appropriate cell. First, if the Auto Calc is on, click on the task bar to turn it off. Otherwise, the program will recalculate between each entry and it will take several seconds for each recalculation. When you click on each input cell, the descriptor becomes highlighted; type information in highlighted cells. It may be necessary to sweep across the value with your mouse or triple-click on the cell before entering the new number; or you can use the up and down arrows for minor changes. All entries are in metric units; conversions and conversion factors are available by clicking on the Conversion option, which displays information and converted equivalents in English units. Enter the energy concentration of your diet. Let’s assume that you are feeding a diet with 3,400 kcal of DE per kg; enter 3400. Daily energy intake will be calculated by the model. If there is a value in this box, then go to the lower part of the screen and find Enter Dietary Energy Intake and click on the ‘‘no’’ box. The energy intake entry box will now be blank. Let’s assume that the sow weighs approximately 175 kg after farrowing and that we expect it to lose approximately 5 kg of weight during a 21-day lactation period. The sow

167

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

168

Nutrient Requirements of Swine

TABLE 1

Growth Model Report Printout.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 4: User’s Guide TABLE 2

Gestation Model Report Printout.

Copyright © National Academy of Sciences. All rights reserved.

169

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

170

Nutrient Requirements of Swine

TABLE 3

Lactation Model Report Printout.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 5 GROWTH MODEL HELP SCREEN

Help Screens

down or by triple-clicking on the input box. Insert the new value from the keyboard, then press ‘‘Enter’’ on the keyboard. The next cell will then be highlighted and a new value can be entered. After all desired values are entered, click on the red ‘‘Re-Calc’’ button. Another way is to click on the up and down arrows inside the input box. If you want to change the increments, press the ‘‘Min/Max’’ button (see Min/Max section for further instructions). Decimals can be changed in the input boxes and on the outputs in the visual and printed reports. Input boxes—highlight the box by clicking on it, then click on the ‘‘Min/Max’’ button (see Min/ Max section for further details). Outputs—click on the output value and a screen will appear allowing you to change the number of decimal points.

Index Tabs for Gestation, Lactation, and Growth Inputs and Decimals Scrolling Dietary Energy Concentration Daily Energy Intake Pig Weight Number of Gilts, Barrows, and Boars Lean Gain Lean Gain Curve Space/Pig Temperature Report Based On Reports Report Buttons A, B, C, D, E, F, G, and Report Heading Min/Max Conversions Auto Calc Off and Auto Calc On Reset Feed Tables Help PreView Print Exit

Scrolling To scroll the report screens, use the scroll slide on the right of the report or click on the variable name and use the up and down arrows on the keyboard. Dietary Energy Concentration Enter the digestible energy (DE) or metabolizable energy (ME) concentration of the diet. DE or ME would have been selected on the opening screen. If you want to change from DE to ME (or from ME to DE), press ‘‘Exit,’’ select ‘‘DE’’ or ‘‘ME,’’ then return to the desired model.

GROWTH MODEL Tabs for Gestation, Lactation, and Growth Click on these tabs to go between models.

Daily Energy Intake Functions only when the radio button in the box at the lower left corner marked ‘‘Enter Dietary Energy Intake’’ is on ‘‘Yes.’’ When the button is on ‘‘No,’’ this box will be blank. If the ‘‘Yes’’ button is on, then enter the daily DE or ME intake.

Inputs and Decimals Inputs can be made in two ways, as follows: One way is to highlight the input box by clicking the box with the mouse and sweeping across the value in the box with the mouse button held

171

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

172

Nutrient Requirements of Swine

Pig Weight Enter pig weight in kg. If pig weight is below 20 kg, the number of pigs in each gender group and the lean gain data will disappear. Number of Gilts, Barrows, and Boars Enter the number of each gender. If you want requirements for barrows only, place a ‘‘1’’ in the box beside ‘‘Barrows’’ and a ‘‘0’’ in the other boxes. If you want requirements for an average of barrows and gilts, place a ‘‘1’’ in the box beside ‘‘Barrows’’ and a ‘‘1’’ in the box beside ‘‘Gilts’’ and a ‘‘0’’ in the box beside ‘‘Boars.’’ Note that these entries do not apply to pigs less than 20 kg body weight. Lean Gain Place the mean carcass fat-free lean gain in the appropriate box beside each gender. The lean gain is the average over the live weight range of 20 to 120 kg. It is determined by taking the fat-free lean in the carcass at slaughter, subtracting the estimated fat-free lean in the carcass at the start, and dividing the difference by the number of days from start to finish (see Appendix 2). Lean Gain Curve Select the desired radio button. The standard button marked ‘‘Std’’ uses the model’s default lean growth curve, and this is ordinarily the one to use. However, if you have information on your herd indicating that the lean growth pattern is different from the one used in the model, your curve can be entered as a quadratic or cubic equation. You will need to use a spreadsheet program, such as illustrated in Appendix 3, to generate the coefficients for a quadratic or cubic equation. To enter different equations, click on ‘‘User.’’ A screen will appear and you can enter the coefficients. If you want to use a quadratic equation, enter a ‘‘0’’ as the fourth coefficient. The new equations can be named and saved for future use. Up to 99 equations can be saved. Click on the ‘‘Return’’ button to return to the main program. Space/Pig Click on the button if you want to enter the stocking density of the pigs, and enter the space per pig in square meters (1 square meter 4 10.76 square ft). The model will then reduce energy intake if pigs are crowded. The adjustments in feed intake are not very precise, so use with caution. If you do not check the button, adjustments are not made for crowding. Temperature Click on the button if you want to enter the average ambient temperature (in centigrade) over a 24-hour period. The model will adjust energy intake upwards

when pigs are at temperatures below their comfort (thermoneutral) zone and will adjust it downward when they are at temperatures above their comfort zone. The ideal temperature is approximately 26°C for 10 kg pigs, 25°C for 20 kg pigs, 23°C for 50 kg pigs, and 20°C for 100 kg pigs. Each 1° deviation from the ideal temperature results in an approximate adjustment in daily DE intake of 18, 75, 125, and 175 kcal of DE for pigs weighing 5, 20, 50, and 100 kg body weight, respectively. If you do not check the ‘‘Temperature’’ button, adjustments in energy intake will not be made. Report Based On Click on ‘‘% or amount/kg’’ if you want the report to give requirements on a dietary concentration basis (% of diet, mg/kg of diet, etc.). Click on ‘‘Amount/d’’ if you want the report to give requirements on a daily basis (g/d, mg/d, etc.). Reports Click on this button to give a screen report of the data as entered and the nutrient requirements. Report Buttons A, B, C, D, E, F, G, and Report Heading Click on button ‘‘A’’ for a report of the data as entered. Body weight—gives body weight as entered. Space/pig—gives space/pig as entered. If space/ pig was not entered, it will show ‘‘****.’’ Temperature—gives temperature as entered. If temperature was not entered, it will show ‘‘****.’’ Carcass lean gain, 20-120 kg—gives the carcass fatfree lean accretion rate as entered. Carcass lean tissue gain—gives the carcass fat-free lean tissue accretion rate at the particular weight that you have chosen, based on the lean growth curve. Whole body protein gain—gives the estimated whole body protein accretion rate at the particular weight that you have chosen, assuming that 2.55 grams of carcass fat-free lean is equivalent to 1 gram of whole body protein. Whole body protein tissue gain—gives the whole body lean tissue accretion rate at the particular weight that you have chosen. Assumes that whole body lean tissue is 23% protein. Whole body fat tissue gain—gives the whole body fat tissue accretion rate at the particular weight that you have chosen. Assumes that whole body fat tissue is 90% fat. DE concentration of diet—gives the DE concentration of the diet as entered. ME concentration of diet—gives the ME concentration of the diet assuming that ME is 96% of DE.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 5: Help Screens DE intake—gives the DE intake based on the default curves in the model. If you entered DE intake, then it gives DE intake as entered. ME intake—gives the ME intake assuming that ME is 96% of DE. Expected daily feed intake—gives the daily feed intake based on the DE intake divided by the DE concentration of the diet. Expected daily gain—gives the expected daily gain of the pigs. This is determined by adding the daily accretion of total body protein tissue and total body fat tissue in the body and dividing by 0.94 (which assumes that 6% of the body weight is gut fill). Protein tissue is dry protein tissue divided by 0.23 (assumes that protein is 77% water) and fat tissue is dry fat tissue divided by 0.90 (assumes that fat tissue is 10% water). Expected feed efficiency—gives the expected feed/ gain by dividing daily feed intake by daily gain. Crude protein (corn–soy diet)—gives the crude protein in a mixture of corn and soybean meal that meets the lysine requirement. It is based on the following relationship: crude protein (%) 4 5.22 ` (15.51 2 true digestible lysine, %) Click on button ‘‘B’’ for the true ileal digestible amino acid requirements. Click on button ‘‘C’’ for the apparent ileal digestible amino acid requirements. Click on button ‘‘D’’ for the total amino acid requirements, assuming that the diet consists of a mixture of corn and soybean meal. Click on button ‘‘E’’ for the mineral requirements. Click on button ‘‘F’’ for the vitamin requirements. Click on button ‘ ‘G ’’ fo r t he es se nt ia l f at ty ac id requirements. Click on ‘‘Report Heading’’ to make a heading for your report. Min/Max Click on this button to give parameter settings for the input boxes. Click on one of the input boxes. The screen now shows the name of the input. Set the minimum and maximum values for that input by pressing the up or down arrows, or by entering the values from the keyboard. Click on ‘‘Save’’ to save those values. Set the increment for that input by choosing numbers in the two boxes marked ‘‘ by the .’’ Insert different numbers and power of you will see how the increment changes. Click on ‘‘Save’’ to save the new increments. Set the desired number of decimal points for the input. Click on ‘‘Save’’ to save the decimal setting.

173

Click on any other input boxes to make desired settings. When finished, click on the ‘‘Report’’ button. Conversions Click on this button to give conversions from metric to standard U.S. equivalents. Click on one of the input boxes. The conversions are shown. When finished, click on the ‘‘Report’’ box. Auto Calc Off and Auto Calc On Click on this button to toggle between off and on. When the ‘‘Auto Calc’’ button is off, this button will change to ‘‘Re Calc Grow-Finish Values’’ after any changes are made in inputs. The red ‘‘Re Calc’’ button must be pressed to give the new requirement values in the report. When the ‘‘Auto Calc’’ button is on, the recalculations will occur automatically after any input is changed. However, this can be time consuming when several changes are made or when the up and down arrows are used to increment between values. Reset Click on this button to reset the model to the default values that you found when the program was opened. Feed Tables Click on this button to access the feed composition tables. Help Click on this button to display the help screen. PreView Click on this button to preview the printed report. Print Click on this button to print the report. If you want to make a heading for your report, press ‘‘Report Heading’’ and type a heading. The report will give the same information that is shown in the screen report. The upper portion shows your inputs and the calculations of various outputs and the lower portion gives the requirements on a diet concentration and daily basis. If the outputs are not appropriate for the particular pig weight chosen or if temperature or space/pig was not an input, the output will display ‘‘****.’’ Exit Click to exit the program.

GESTATION MODEL HELP SCREEN Index Tabs for Gestation, Lactation, and Growth Inputs and Decimals

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

174

Nutrient Requirements of Swine

Scrolling Dietary Energy Concentration Daily Energy Intake Sow Weight at Breeding Gestation Weight Gain Number Pigs Born Temperature Report Based On Reports Report Buttons A, B, C, D, E, F, G, and Report Heading Min/Max Conversions Auto Calc Off and Auto Calc On Reset Feed Table Help PreView Print Exit

Dietary Energy Concentration Enter the digestible energy (DE) or metabolizable energy (ME) concentration of the diet. DE or ME would have been selected on the opening screen. If you want to change from DE to ME (or from ME to DE), press ‘‘Exit,’’ select ‘‘DE’’ or ‘‘ME,’’ then return to the desired model. Daily Energy Intake Functions only when the radio button in the box at the lower left corner marked ‘‘Enter Dietary Energy Intake’’ is on ‘‘Yes.’’ When the button is on ‘‘No,’’ this box will be blank. If the ‘‘Yes’’ button is on, then enter the daily DE or ME intake. Sow Weight at Breeding Enter the breeding weight of the sow in kg.

GESTATION MODEL

Gestation Weight Gain Enter the desired weight gain of the sow during gestation. The weight gain includes both the sow body and the products of conception. Note that an entry cannot be made in this box if DE or ME intake is an input. When that is the case, gestation weight gain will be calculated by the model.

Tabs for Gestation, Lactation, and Growth Click on these tabs to go between models.

Litter Size, Total Pigs Enter the anticipated size of the litter, total pigs.

Inputs and Decimals Inputs can be made in two ways, as follows: One way is to highlight the input box by clicking the box with the mouse and sweeping across the value in the box with the mouse button held down or by triple-clicking on the input box. Insert the new value from the keyboard, then press ‘‘Enter’’ on the keyboard. The next cell will then be highlighted and a new value can be entered. After all desired values are entered, click on the red ‘‘Re-Calc’’ button. Another way is to click on the up and down arrows inside the input box. If you want to change the increments, press the ‘‘Min/Max’’ button (see Min/Max section for further instructions). Decimals can be changed in the input boxes and on the outputs in the visual and printed reports. Input boxes—highlight the box by clicking on it, then click on the ‘‘Min/Max’’ button (see Min/Max section for further details). Outputs—click on the output value and a screen will appear allowing you to change the number of decimal points.

Temperature Click on the button if you want to enter the average ambient temperature (in centigrade) over a 24-hour period. The model will adjust energy intake or gestation weight gain when the temperature is below 20°C. Approximately 230 kcal of ME per day is required for each 1° below 20°C. No adjustment is made for temperatures above 20°C. If you do not check the ‘‘Temperature’’ button, an adjustment will not be made.

Scrolling To scroll the report screens, use the scroll slide on the right of the report or click on the variable name and use the up and down arrows on the keyboard.

Report Based On Click on ‘‘% or amount/kg’’ if you want the report to give requirements on a dietary concentration basis (% of diet, mg/kg of diet, etc.). Click on ‘‘Amount/d’’ if you want the report to give requirements on a daily basis (g/d, mg/d, etc.). Reports Click on this button to give a screen report of the data as entered and the nutrient requirements. Report Buttons A, B, C, D, E, F, G, and Report Heading Click on button ‘‘A’’ for a report of the data as entered. Sow weight at breeding—gives the breeding weight as entered. Expected gestation weight gain—gives the gestation weight gain as entered; or gives it as calculated, if DE intake was entered.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 5: Help Screens Expected litter size—gives litter size, total pigs, as entered. Conceptus gain—gives the estimated weight gain of the products of conception, including the pigs, placenta, and placental fluids during the entire gestation period (2.28 2 number of pigs). Maternal gain—gives the estimated weight gain of maternal tissue excluding the litter during the entire gestation period. Maternal lean tissue gain—gives the estimated weight gain of maternal lean tissue during the entire gestation period. Maternal fat tissue gain—gives the estimated weight gain of maternal fat tissue during the entire gestation period. Temperature—gives temperature as entered. If temperature was not entered, it will show ‘‘****.’’ DE concentration of diet—gives the DE concentration of the diet as entered. ME concentration of diet—gives the ME concentration of the diet assuming that ME is 96% of DE. DE intake—gives DE intake as entered; or gives it as calculated if gestation weight gain was entered. ME intake—gives the ME intake assuming that ME is 96% of DE. Expected daily feed intake—gives the daily feed intake based on the DE intake divided by the DE concentration of the diet. Crude protein (corn–soy diet)—gives the crude protein in a mixture of corn and soybean meal that meets the lysine requirement. It is based on the following relationship: crude protein (%) 4 5.22 ` (15.51 2 true digestible lysine, %) Click on button ‘‘B’’ for the true ileal digestible amino acid requirements. Click on button ‘‘C’’ for the apparent ileal digestible amino acid requirements. Click on button ‘‘D’’ for the total amino acid requirements, assuming that the diet consists of a mixture of corn and soybean meal. Click on button ‘‘E’’ for the mineral requirements. Click on button ‘‘F’’ for the vitamin requirements. Click on button ‘‘G’’ for the essential fatty acid requirements. Click on ‘‘Report Heading’’ to make a heading for your report. Min/Max Click on this button to give parameter settings for the input boxes.

175

Click on one of the input boxes. The screen now shows the name of the input. Set the minimum and maximum values for that input by pressing the up or down arrows, or by entering the values from the keyboard. Click on ‘‘Save’’ to save those values. Set the increment for that input by choosing numby the bers in the two boxes marked ‘‘ power of .’’ Insert different numbers and you will see how the increment changes. Click on ‘‘Save’’ to save the new increments. Set the desired number of decimal points for the input. Click on ‘‘Save’’ to save the decimal setting. Click on any other input boxes to make desired settings. When finished, click on the ‘‘Report’’ button. Conversions Click on this button to give conversions from metric to standard U.S. equivalents. Click on one of the input boxes. The conversions are shown. When finished, click on the ‘‘Report’’ box. Auto Calc Off and Auto Calc On Click on this button to toggle between off and on. When the ‘‘Auto Calc’’ button is off, this button will change to ‘‘Re-Calc Gestation Values’’ after any changes are made in inputs. The red ‘‘ReCalc’’ button must be pressed to give the new requirement values in the report. When the ‘‘Auto Calc’’ button is on, the recalculations will occur automatically after any input is changed. However, this can be time consuming when several changes are made or when the up and down arrows are used to increment between values. Reset Click on this button to reset the model to the default values that you found when the program was opened. Feed Tables Click on this button to access the feed composition tables. Help Click on this button to display the help screen. PreView Click on this button to preview the printed report. Print Click on this button to print the report. If you want to make a heading for your report, press ‘‘Report Heading’’ and type a heading. The report will give the

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

176

Nutrient Requirements of Swine

same information that is shown in the screen report. The upper portion shows your inputs and the calculations of various outputs and the lower portion gives the requirements on a diet concentration and daily basis. If temperature was not an input, the output will display ‘‘****.’’ Exit Click to exit the program.

LACTATION MODEL HELP SCREEN Index Tabs for Gestation, Lactation, and Growth Inputs and Decimals Scrolling Dietary Energy Concentration Daily Energy Intake Sow Weight after Farrowing Lactation Length Lactation Weight Change Number Pigs Nursed Daily Pig Weight Gain Temperature Report Based On Reports Report Buttons A, B, C, D, E, F, G, and Report Heading Min/Max Conversions Auto Calc Off and Auto Calc On Reset Feed Tables Help PreView Print Exit

LACTATION MODEL Tabs for Gestation, Lactation, and Growth Click on these tabs to go between models. Inputs and Decimals Inputs can be made in two ways, as follows: One way is to highlight the input box by clicking the box with the mouse and sweeping across the value in the box with the mouse button held down or by triple-clicking on the input box. Insert the new value from the keyboard, then press ‘‘Enter’’ on the keyboard. The next cell will then be highlighted and a new value

can be entered. After all desired values are entered, click on the red ‘‘Re-Calc’’ button. Another way is to click on the up and down arrows inside the input box. If you want to change the increments, press the ‘‘Min/Max’’ button (see Min/Max section for further instructions). Decimals can be changed in the input boxes and on the outputs in the visual and printed reports. Input boxes—highlight the box by clicking on it, then click on the ‘‘Min/Max’’ button (see Min/ Max section for further details). Outputs—click on the output value and a screen will appear allowing you to change the number of decimal points. Scrolling To scroll the report screens, use the scroll slide on the right of the report or click on the variable name and use the up and down arrows on the keyboard. Dietary Energy Concentration Enter the digestible energy (DE) or metabolizable energy (ME) concentration of the diet. DE or ME would have been selected on the opening screen. If you want to change from DE to ME (or from ME to DE), press ‘‘Exit,’’ select ‘‘DE’’ or ‘‘ME,’’ then return to the desired model. Daily Energy Intake Functions only when the radio button in the box at the lower left corner marked ‘‘Enter Dietary Energy Intake’’ is on ‘‘Yes.’’ When the button is on ‘‘No,’’ this box will be blank. If the ‘‘Yes’’ button is on, then enter the daily DE or ME intake. Sow Weight after Farrowing Enter the postfarrowing weight of the sow in kg. Lactation Length Enter the number of days that the sow will nurse the litter. Lactation Weight Change Enter the desired or anticipated weight change from the beginning to the end of lactation. If no weight change is targeted, then enter ‘‘0.’’ Weight loss is entered as a negative value. Note that an entry cannot be made in this box if DE or ME intake is an input. When that is the case, lactation weight change will be calculated by the model. Number Pigs Nursed Enter the number of pigs nursed by the sow during the lactation period. Daily Pig Weight Gain Enter the average daily weight gain of the pigs during the lactation period.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Appendix 5: Help Screens Temperature Click on the button if you want to enter the average farrowing room temperature (in centigrade) over a 24hour period. The model will adjust energy intake or lactation weight change when the temperature deviates from 20°C. An additional 310 kcal of ME per day is required per 1° below 20°C, and 310 fewer kcal of ME per day are required for each 1° above 20°C. If you do not check the ‘‘Temperature’’ button, adjustments will not be made. Report Based On Click on ‘‘% or amount/kg’’ if you want the report to give requirements on a dietary concentration basis (% of diet, mg/kg of diet, etc.). Click on ‘‘Amount/d’’ if you want the report to give requirements on a daily basis (g/d, mg/d, etc.). Reports Click on this button to give a screen report of the data as entered and the nutrient requirements. Report Buttons A, B, C, D, E, F, G, and Report Heading Click on button ‘‘A’’ for a report of the data as entered. Sow weight at farrowing—gives the sow’s postfarrowing weight as entered. Expected lactation weight change—gives the lactation weight change as entered; or gives it as calculated, if DE intake was entered. Lactation length—gives the length of lactation as entered. No. pigs nursed—gives the number of pigs nursed during the lactation period as entered. Avg daily gain of nursed pigs—gives the average daily gain per pig over the lactation period, as entered. Farrowing room temperature—gives the temperature as entered. If temperature was not entered, it will show ‘‘****.’’ Maternal lean tissue gain—gives the estimated maternal lean tissue weight gain or loss during the lactation period. Maternal fat tissue gain—gives the estimated maternal fat tissue weight gain or loss during the lactation period. Estimated milk production—gives the estimated daily milk production based on the weight gain of the litter. DE concentration of diet—gives the DE concentration of the diet as entered. ME concentration of diet—gives the ME concentration of the diet assuming that ME is 96% of DE. DE intake—gives DE intake as entered; or gives it as calculated if lactation weight change was entered.

177

ME intake—gives the ME intake assuming that ME is 96% of DE. Expected daily feed intake—gives the daily feed intake based on the DE intake divided by the DE concentration of the diet. Crude protein (corn–soy diet)—gives the crude protein in a mixture of corn and soybean meal that meets the lysine requirement. It is based on the following relationship: crude protein (%) 4 5.22 ` (15.51 2 true digestible lysine, %) Click on button ‘‘B’’ for the true ileal digestible amino acid requirements. Click on button ‘‘C’’ for the apparent ileal digestible amino acid requirements. Click on button ‘‘D’’ for the total amino acid requirements, assuming that the diet consists of a mixture of corn and soybean meal. Click on button ‘‘E’’ for the mineral requirements. Click on button ‘‘F’’ for the vitamin requirements. Click on button ‘‘G’’ for the essential fatty acid requirements. Click on ‘‘Report Heading’’ to make a heading for your report. Min/Max Click on this button to give parameter settings for the input boxes. Click on one of the input boxes. The screen now shows the name of the input. Set the minimum and maximum values for that input by pressing the up or down arrows, or by entering the values from the keyboard. Click on ‘‘Save’’ to save those values. Set the increment for that input by choosing numby the bers in the two boxes marked ‘‘ .’’ Insert different numbers and power of you will see how the increment changes. Click on ‘‘Save’’ to save the new increments. Set the desired number of decimal points for the input. Click on ‘‘Save’’ to save the decimal setting. Click on any other input boxes to make desired settings. When finished, click on the ‘‘Report’’ button. Conversions Click on this button to give conversions from metric to standard U.S. equivalents. Click on one of the input boxes. The conversions are shown. When finished, click on the ‘‘Report’’ box. Auto Calc Off and Auto Calc On Click on this button to toggle between off and on.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

178

Nutrient Requirements of Swine When the ‘‘Auto Calc’’ button is off, this button will change to ‘‘Re-Calc Gestation Values’’ after any changes are made in inputs. The red ‘‘ReCalc’’ button must be pressed to give the new requirement values in the report. When the ‘‘Auto Calc’’ button is on, the recalculations will occur automatically after any input is changed. However, this can be time consuming when several changes are made or when the up and down arrows are used to increment between values.

Reset Click on this button to reset the model to the default values that you found when the program was opened. Feed Tables Click on this button to access the feed composition tables.

Help Click on this button to display the help screen. PreView Click on this button to preview the printed report. Print Click on this button to print the report. If you want to make a heading for your report, press ‘‘Report Heading’’ and type a heading. The report will give the same information that is shown in the screen report. The upper portion shows your inputs and the calculations of various outputs and the lower portion gives the requirements on a diet concentration and daily basis. If temperature was not an input, the output will display ‘‘****.’’ Exit Click to exit the program.

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Authors

Gary L. Cromwell (Chair) is a professor of animal science at the University of Kentucky at Lexington. He received his Ph.D. from Purdue University. His research interests include mineral and amino acid nutrition of swine, and efficacy and safety of feed additives.

Austin J. Lewis is a professor of animal nutrition at the University of Nebraska. He received his Ph.D. from the University of Nottingham, in the United Kingdom. His research interests include nutritional requirements of swine, especially proteins and amino acids.

David H. Baker serves as professor of animal nutrition at the University of Illinois at Urbana–Champaign. He received his Ph.D. in nutrition with a microbiology and biochemistry minor from the University of Illinois. His research areas include amino acid nutrition and metabolism.

James E. Pettigrew, a professor of animal science at the University of Minnesota until 1997, recently began his own consulting business, Pettigrew Consulting International. He received his Ph.D. from the University of Illinois. Research interests include swine growth and production, modeling, reproduction, and lactation. Norman C. Steele is research leader for the USDA/ARS Growth Biology Laboratory in Beltsville, Maryland. He received his Ph.D. in Dairy Science from the University of Maryland. Research interests include maternal-fetal nutrient interactions, the effect of dietary energy on protein utilization, and management factors affecting efficacy of metabolism modifiers.

Richard C. Ewan is a professor of animal science at Iowa State University. He received his Ph.D. from the University of Wisconsin. Research interests include swine growth, genetic background and selenium status in swine, and nutrient bioavailability in swine. E. T. Kornegay is a professor of animal science at Virginia Polytechnic Institute and State University. He received his Ph.D. from Michigan State University. Major areas of research include environmental nutrition, sow and piglet management and nutrition, mineral availability of inorganic and organic sources, and evaluation of feedstuffs for swine.

Philip A. Thacker is currently professor of animal science at the University of Saskatchewan, Canada. He received his Ph.D. from the University of Alberta. Research interests include swine growth, nontraditional feed sources for use in swine production, and nutrient metabolism in swine.

179

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Index

A

growing-finishing pigs, requirements of, 35 ileal digestibilities, 18–19, 35–36 isomers, 19 lactation requirements, 38–40 model for predicting requirements, 31–32 nutrient requirement tables, 110 ratios, 17–18 synthesis, 16 weanling pig requirements, 40–41 See also specific acid Anemia, iron-deficiency, 54–55 Anthelmintic additives, 98 Antibiotics, 97, 99–100 water intake and, 93 Antimicrobial additives growth effects, 97–98 physiological function, 97 safety concerns, 99–100 Antioxidants, 74, 99 Arachidonic acid, 9 Arginine, 16, 17 deficiencies and excesses, 19 Ascorbic acid. See Vitamin C Avidin, 76

Acid-base status, 50 Acid detergent fiber, 8 composition of feed ingredients, 124 Acidifiers, 98 Activity, 5 Additives, feed acidifiers in, 98 anthelmintic, 98 antimicrobial, 97–98 antioxidants, 99 carcass modifiers, 99 enzyme, 98 flavor, 99 flow agents, 99 microbial supplements, 98 mineral supplements, 99 odor control agents, 99 oligosaccharide, 98 pellet binders, 99 regulations, 100 safety concerns, 99–100 Aflatoxin, 99 Alfalfa calcium bioavailability, 49 phosphorus content, 49 Amino acids antagonisms, 19 bioavailability, 18–19 composition of feed ingredients, 124–125 deficiencies and excesses, 19 dietary requirements, 17, 19–25 essential/nonessential, 16–17 excretory losses of nutrient and, 104 feed analysis, 17 gestation requirements, 38

B B vitamins, 71. See also specific B vitamin Barley, pantothenic acid in, 78 Barrows, nutrient requirements of, 110 Bioavailability amino acids, 18–19 biotin, 75 calcium, 49 choline, 76 copper, 53 iron, 54

181

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

182

Nutrient Requirements of Swine

phosphorus, 48–49 vitamin A, 72 vitamin B12, 80–81 vitamin C, 81 Biotin bioavailability, 75 deficiency, 76 requirements, 76 Birth weights, 7 copper effects, 53 manganese effects, 55 vitamin A requirements, 72 vitamin B12 effects, 81 Boars amino acid requirements, 25–26 energy requirements, 7–8 maintenance requirement, 6 mineral requirements, 48 water requirements, 93 zinc requirement, 56 See also Breeding boars Body protein mass, maintenance requirements, 6 Body temperature energy metabolism, 5 See also Thermoregulation Body weight, metabolic, 6 Bone growth calcium-phosphorus deficiency, 49 manganese requirements, 55 mineral requirements, 47, 48, 50 vitamin D requirements, 73 Breeding boars amino acid requirements, 25–26 chlorine requirements, 50 energy requirements, 7–8 magnesium requirements, 50 mineral requirements, 48 nutrient requirement tables, 110 semen quality, 25–26 sodium requirements, 50 zinc requirement, 57 C Calcitonin, 73 Calcium, 47 bioavailability, 49 deficiencies/excesses, 49 excretion, 103, 104 phosphorous ratio, 47, 49 requirements, 47–48, 49 vitamin D interaction, 73 Carbohydrates, gross energy, 3 b-Carotene, 2, 72

Carotenoids, 71, 72 Cereal grains calcium bioavailability, 49 phosphorus content, 48 See also specific grain Chemical composition feed intake and, 10 net energy and, 5 predicting digestible energy from, 3–4 Chloride-potassium interaction, 51 Chlorine, 47 deficiencies and excesses, 50 requirements, 49–50 Cholecalciferol, 73 Choline bioavailability, 76 deficiency, 77 forms of, 76 function, 76–77 requirement, 77 toxicity, 77 Chromium, 47 absorption, 51 metabolic function, 51, 52 picolinate, 51, 52 tolerance, 52 Cobalt, 47, 52 Copper, 47, 99 bioavailability, 53 deficiencies and excesses, 53 excretion, 103–104 growth effects, 52, 53 requirements, 52 Corn niacin availability, 78 pantothenic acid availability, 78 phosphorus content, 48 riboflavin availability, 79 vitamin B6 availability, 80 Corn–soybean meal diet, 105–109 Crude fiber, 8 Cyanocobalamin. See Vitamin B12 Cysteine, 16–17 methionine and, 16–17 Cystine requirements, 24 D Defecation water loss, 91 See also Excretion Deficiencies amino acid, 19

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Index arginine, 19 biotin, 76 calcium, 49 chlorine, 50 choline, 77 copper, 53 folacin, 77–78 iodine, 53 iron, 54–55 magnesium, 50 niacin, 78 pantothenic acid, 79 phosphorus, 49 potassium, 51 riboflavin, 79 selenium, 56 sodium, 50 thiamin, 80 vitamin A, 72 vitamin B6, 80 vitamin B12, 81 vitamin D, 73 vitamin E, 74 vitamin K, 75 Diarrhea, 19 Dichlorvos, 98 Diet formulation amino acids in, 17 density, 16–17 manure management and, 105 nutrient composition, 124 procedures, 105–109 Digestibility amino acid bioavailability, 18–19 apparent ileal, 18–19, 32–40 enzyme additives for, 98 fat, 9–10 fiber and, 8 true ileal, 18–19, 32–40 Digestible energy (DE) calculations, 3–4 fiber content of feed and, 8 growing-finishing pigs, requirements of, 33–34 true/apparent, 3 E Electrolyte balance, 50, 51 Energy classification, 3–5 digestible (DE), 3–4 gross, 3 heat production, 5 metabolizable (ME), 4

net (NE), 4–5 nutrient requirement tables, 110 optimal protein-to-energy ratio, 9 terminology, 3 Energy requirements boars and gilts, 7–8 gestation, 6–7, 36–38 growing-finishing pigs, 33–34 growth, 6 lactation, 6, 7, 38–39 maintenance, 5–6 model for estimating, 31–32 reproduction, 8 weanling pigs, 40–41 Energy sources lipids, 9–10 nonstarch polysaccharides, 8–9 starch, 8 sugars, 8 Environmental temperature, 5 diet and, 8 water intake and, 92–93 water loss and, 91 Ergocalciferol, 73 Essential fatty acid requirements, 9 Excretion minimizing nutrient loss, 104–105 nutrient losses, 103–104 pollution concerns, 103 F Fat digestibility, 9–10 energy costs of protein retention, 6 gestation requirements, 37 gross energy, 3 lipids, 9–10 nutrient composition of feed, 125 Fatty acids, 2 digestibility, 9–10 essential, 9 volatile, 8 Feed intake determinants of, 10 developmental requirements, 10–11 environmental conditions and, 5, 10 litter size and birth weight, 6–7 water intake and, 92, 93 weanling pig, 41 Feedstuffs additives. See Additives, feed amino acid bioavailability, 18–19 amino acid requirements, 17

Copyright © National Academy of Sciences. All rights reserved.

183

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

184

Nutrient Requirements of Swine

calcium-phosphorus content, 47–49 energy classification, 3–5 expression of energy content, 3 fiber content, 8 nutrient composition, 124–125 vitamin A stability, 72–73 vitamin E availability, 73–74 vitamin K stability, 75 Fiber digestibility and, 8–9 measuring feed content, 8 Flavoring agents, 99 Flow agents, 99 Folacin deficiency, 77–78, 81 function, 77 requirement, 77 Fructose, 8 G Gestation amino acid requirements, 25, 38 biotin effects, 76 calcium-phosphorus requirements, 48 chlorine requirements, 50 copper requirements, 52, 53 dietary fat during, 10 energy requirements, 6–7, 36–38 feed intake, 6–7, 11 folacin effects, 77 iodine intake, 54 iron levels, 54 magnesium requirements, 50 maintenance requirements, 37, 38 manganese requirements, 55 nutrient requirement modeling, 44 nutrient requirement tables, 110 riboflavin requirement, 79 selenium effects, 55 sodium requirements, 50 thermoregulation in, 37–38 vitamin A requirements, 72 vitamin B6 requirement, 80 vitamin C effects, 81 vitamin D requirements, 73 vitamin E requirements, 74 water requirements, 93 weight gain, 6, 7, 36–37 zinc requirement, 57 See also Birth weights; Litter size Gilts energy requirements, 7–8 mineral requirements, 48

nutrient requirement tables, 110 zinc requirement, 56 Glucose, 8 chromium and, 51 Glutamine, 17 Glutathione peroxidase, 56 Gross energy (GE), 3 Growing-finishing pigs amino acid requirements, 17, 24–25, 32–36 biotin effects, 76 calcium-phosphorus requirements, 47–49 chlorine requirements, 50 energy requirements, 33–34 feed intake, 10 magnesium requirements, 50 niacin requirements, 78 nutrient requirement tables, 110 selenium requirement, 55 sodium requirements, 50 thiamin requirements, 79 vitamin A requirements, 72 water requirements, 92–93 Growth antimicrobial effects, 97–98 biotin effects, 76 boars and gilts, 7 calcium requirements, 47–49 cobalt and, 52, 53 copper effects, 52, 53 energy requirements, 6 energy sources for, 8–10 folacin effects, 77–78 iron requirements, 54 lipid requirements, 9 manganese requirements, 55 nutrient requirement variables, 31 phosphorus requirements, 47–49 potassium requirements, 51 riboflavin requirement, 79 thiamin requirement, 79 vitamin A requirements, 72 water requirements, 92 zinc requirement, 56, 57

H Heat increment (HI), 4 Heat production, 5 Histidine, 17 requirements, 24 Hypervitaminosis vitamin A, 72 vitamin D, 73

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Index I Ideal protein, 17 Ileal digestibilities, 18–19 amino acid requirements, 35–36 composition of feed ingredients, 124 lysine, 34–35 true/apparent, 18–19, 124 Infectious organisms in water, 93–95 Intake. See Feed intake Iodine, 47, 53–54 Iron, 47 bioavailability, 54 deficiencies, 54–55 excretion, 103–104 function, 54 hemoglobin levels, 54 niacin availability and, 78 requirements, 54 toxicity, 55 Ivermectin, 98 L Lactation amino acid ratios, 17 amino acid requirements, 25, 38, 39–40 antimicrobial effects, 97 biotin effects, 76 calcium-phosphorus requirements, 48 chlorine requirements, 50 choline effects, 77 copper effects, 53 dietary fat during, 10 energy requirements, 6, 7, 38–39 feed intake, 11 folacin effects, 77 iodine intake, 54 iron levels, 54 magnesium requirements, 50 manganese requirements, 55 nutrient requirements, 38, 44, 110 riboflavin requirements, 79 sodium requirements, 50 vitamin B6 requirements, 80 vitamin B12 requirements, 81 vitamin D requirements, 73 vitamin E requirements, 74 water requirements, 93 zinc requirements, 57 Lactose, 8 Leucine, 19 Linoleic acid, 9 requirements, 110 Lipids, 9–10. See also Fat

Litter size, 6–7 choline effects, 76–77 chromium and, 52 folacin effects, 77 vitamin A requirements, 72 vitamin B12 effects, 81 Lysine deficiencies and excesses, 19 gestation requirements, 38 growing-finishing pig requirements, 32–35 growth model, 42–44 isomers, 19 lactation requirements, 39–40 requirements, 19–26 weanling pig requirements, 40–41 M Magnesium, 47 deficiencies and excesses, 50 excretion, 103–104 requirements, 50 Maintenance amino acid ratios, 17 body protein mass and, 6 energy metabolism, 4–5 in gestation, 6–7, 37 growing-finishing pigs, 32 in lactation, 6, 7, 39 metabolizable energy requirements, 5–6 net energy requirements, 5–6 Manganese, 55 excretion, 103–104 Manure health concerns in distribution of, 103 nutrient content, 103 odor control agents, 99 total production management, 105 Mastitis-metritis-agalactia complex, 74 Meat and bone as feed, 49 Menadiones, 75 Menaquinones, 75 Metabolic body weight, 6 Metabolizable energy (ME), 4 gestation requirements, 37–38 for maintenance, 5–6 Methionine, 17 choline and, 77 cysteine and, 16–17 isomers, 19 requirements, 24, 25–26 Microbial supplements, 98 Milk composition fat content, 10

Copyright © National Academy of Sciences. All rights reserved.

185

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

186

Nutrient Requirements of Swine

iron, 54 vitamin E, 74 Milk production energy requirement, 7 lysine requirement, 39–40 protein requirement, 17, 18 See also Lactation Minerals composition of feed ingredients, 125 feed additives, 99 function, 47 predictive modeling, 41–42 requirements, 47, 110 toxicity, 47 See also specific mineral Model evaluation of, 42–44 gestating sow, 36–38 growing-finishing, 32–36 lactating sow, 38–41 principles for development, 31 Modeling, mathematical, 1. See also Appendixes amino acid requirements, 31–35 energy requirements, 32–34 Muscle growth, 6 N Net energy (NE), 4–5 composition of feed ingredients, 124 maintenance requirements, 6 Neutral detergent fiber, 8 composition of feed ingredients, 124 Niacin, 78 Nicotinamide, 78 Nicotinate, 51, 52 Nicotinic acid, 78 Nitrates/nitrites, in water, 94 Nitrogen excretion, 17, 103, 104 metabolizable energy, 4 minimizing excretory losses, 104–105 pollution prevention, 103 urea supplements, 16 Nutrient requirements barrows, 113–114 boars, 123 gestating sows, 117–118, 121–122 gilts, 113–114 growing pigs, 112–113, 115–116 lactating sows, 119–122 O Oats, niacin availability, 78

Oils, 9, 125 Oilseed meal, phosphorus content, 48–49 Oligosaccharide additives, 98 Osteochondrosis, 81 Osteomalacia, 49 P Pantothenic acid, 78–79 Parakeratosis, 57 Paralysis, posterior, 49 Parasite control, 98 Pellet binders, 99 Phenylalanine, 17 and tyrosine, 17–19 Phosphorus, 47 bioavailability, 48–49 calcium ratio, 47, 49 composition of feed ingredients, 124 deficiencies/excesses, 49 excretion, 103, 104 inorganic, 49 in manure, 103 minimizing excretory losses, 104–105 phytate, 49 pollution prevention, 103 requirements, 47–49 sources, 48–49 Phylloquinones, 75 Phytase, 48–49, 98, 104 Phytic acid, 49, 56, 98 Pigs. See Barrows; Boars; Breeding boars; Gilts; Growing-finishing pigs; Sows; Starting pigs; Suckling pigs; Weanling pigs Polysaccharides, nonstarch energy sources, 8–9 Postweaning pig scouring, 57 zinc requirement, 57 Potassium, 47 deficiencies and excesses, 51 dietary function, 51 excretion, 103–104 mineral interactions, 51 requirements, 50, 51 Predicting nutrient requirements amino acids, 31–32 digestible energy, 3–4 minerals, 41–42 vitamins, 41–42 Pregnancy. See Gestation Proline, 16 Protein accretion, 17, 18

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Index body tissue, 18 definition, 16 digestibility, 8–9 energy ratio, 9 excretory losses of nutrient and, 104 gestation requirements, 37, 38 gross energy, 3 in growing-finishing pigs, 32–35 ideal, 17 retention, energy costs of, 6 sources, 17, 49 See also Amino acids Pyridoxal, 80 Pyridoxal phosphate, 80 Pyridoxamine, 80 Pyridoxine, 80 Pyridoxines. See Vitamin B6 R Regulations, 100 Reproduction antimicrobial effects, 97 choline effects, 76–77 pantothenic acid requirements, 79 vitamin B12 requirements, 81 Resistance to antimicrobials, 99–100 Respiration, water loss in, 90 Retinoids, 71–72 Retinol, 71–72 Retinyl palmitate, 72 Riboflavin, 79 Rickets, 49, 73 S Salt, 50 water intake and, 93, 95 Sarsaponin, 99 Selenium, 47, 55–56 Sexual behavior. See Breeding boars; Reproduction Sodium, 47 deficiencies and excesses, 50 excretion, 103–104 potassium interaction, 51 requirements, 49–50 Somatotropin, 48 Sorghum niacin availability, 78 pantothenic acid availability, 78 Sows amino acid requirements, 25 See also Gestation; Lactation Soybean meal

corn-soybean meal diet, 105–109 iron availability, 54 niacin availability, 78 pantothenic acid availability, 78 riboflavin availability, 79 vitamin B6 availability, 80 Starch, 8 Starting pigs amino acid requirements, 19–24 nutrient requirement tables, 110 Suckling pigs creep feed consumption, 10, 91 feed intake, 10 iron requirements, 54 magnesium requirements, 50 selenium requirement, 55 vitamin C requirements, 81 water requirements, 91–92 Sucrose, 8 Sugars, 8 Sulfa drugs, 78 Sulfates, in water, 94 Sulfur, 47, 51 Sulfur amino acids, 16–17, 51 boar requirements, 25 cysteine-methionine, 16–17 Sweating, 91 Swine. See Pigs T Temperature effects energy metabolism, 5 feed intake, 8, 10 moisture loss, 90–91 protein accretion, 34 vitamin E stability, 74 water consumption, 92–93 Thermoregulation in gestation, 37–38 in lactation, 39 Thiamin, 79–80 Thiaminase, 80 Threonine gestation requirements, 38 isomers, 19 requirements, 24 Thyroid disorders, 53 Thyroxine, 53 Tocopherols, 73–74 Tocotrienols, 73 Total dissolved solids, 93–95 Toxicity chromium, 52

Copyright © National Academy of Sciences. All rights reserved.

187

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

188

Nutrient Requirements of Swine

cobalt, 52 copper, 53 iron, 55 magnesium, 50 manganese, 55 minerals, 47 potassium, 51 selenium, 56 sodium, 50 vitamin A, 71, 72 vitamin D, 71, 73 vitamin E, 74 zinc, 57 Trace minerals, 47 See also specific mineral, 47 Tryptophan, 17 isomers, 19 niacin requirements and, 78 requirements, 24 Tyrosine, 17 U Urea, 16 Urination energy losses, 4 water loss, 91 See also Excretion V Valine lactation requirements, 40 requirements, 24 Vitamin A bioavailability, 72 deficiencies, 72 in feeds, 72–73 function, 71 requirements, 72 retinol equivalents, 71–72 toxicity, 72 vitamin E interaction, 74 Vitamin B6, 80 Vitamin B12, 52 bioavailability, 80–81 deficiency, 81 function, 80 requirements, 71, 81 sources, 80 Vitamin C bioavailability, 81 function, 81 requirements, 82

Vitamin D in calcium-phosphorus metabolism, 47 deficiency, 73 forms of, 73 function, 73 metabolism, 73 requirements, 73 toxicity, 73 Vitamin E, 55 composition of feed ingredients, 124 deficiency, 74 forms of, 73–74 function, 74 requirements, 74 toxicity, 74 Vitamin K deficiency, 74–75 forms of, 75 function, 74–75 requirements, 75 stability in feeds, 75 toxicity, 75 Vitamins composition of feed ingredients, 124 fat-soluble, 71 function, 71 nutrient requirement tables, 110 predictive modeling of requirements, 41–42 toxicity, 71 water-soluble, 71 See also specific vitamin Volatile fatty acids, 8 W Water chlorination, 94–95 functions of, 90 hardness, 94–95 nitrate concentration, 94 pH, 94 physiological function, 90 quality, 93–95 requirements, 91–93 research needs, 90 respiration loss, 90 sulfates in, 94 temperature, 92–93 total dissolved solids, 93–94 turnover, 90–91 Weanling pigs amino acid requirements, 40–41 calcium-phosphorus requirements, 47–49 energy requirements, 40–41

Copyright © National Academy of Sciences. All rights reserved.

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Index feed intake, 10 magnesium requirements, 50 nutrient requirement tables, 110 selenium requirement, 55 survival, sow’s fat intake and, 10 water requirements, 91–92 Weight gain in gestating sow, 6, 7, 36–37 vitamin C effects, 81

Wheat niacin availability, 78 pantothenic acid availability, 78 phosphorus content, 48 Z Zeolite, 99 Zinc, 47, 56–57, 99 excretion, 103–104

Copyright © National Academy of Sciences. All rights reserved.

189

Nutrient Requirements of Swine: 10th Revised Edition http://www.nap.edu/catalog/6016.html

Recent Publications of the Board on Agriculture Policy and Resources Forested Landscapes in Perspective: Prospects and Opportunities for Sustainable Management of America’s Nonfederal Forests 264 pp., ISBN 0-309-05641-1 Precision Agriculture in the 21st Century: Geospatial and Information Technologies in Crop Management (1997), 149 pp., ISBN 0-309-05893-7 Wood in Our Future: The Role of Life-Cycle Analysis, (1997), 147 pp., ISBN 0-309-5745-0 Colleges of Agriculture at the Land Grant Universities: Public Service and Public Policy, (1996), 120 pp., ISBN 0-309-05433-8 Ecologically Based Pest Management: New Solutions for a New Century (1996), 152 pp., ISBN 0-309-05330-7 Colleges of Agriculture at the Land Grant Universities: A Profile (1995), 146 pp., ISBN 0-309-05295-5 Investing in the National Research Initiative: An Update of the Competitive Grants Program in the U.S. Department of Agriculture (1994), 66 pp., ISBN 0-309-05235-1 Rangeland Health: New Methods to Classify, Inventory, and Monitor Rangelands (1994), 180 pp., ISBN 0-309-04879-6 Soil and Water Quality: An Agenda for Agriculture (1993), 516 pp., ISBN 0-309-04933-4 Managing Global Genetic Resources: Agricultural Crop Issues and Policies (1993), 450 pp., ISBN 0-309-04430-8 Pesticides in the Diets of Infants and Children (1993), 408 pp., ISBN 0-309-04875-3 Managing Global Genetic Resources: Livestock (1993), 294 pp., ISBN 0-309-04394-8 Sustainable Agriculture and the Environment in the Humid Tropics (1993), 720 pp., ISBN 0-309-04749-8 Agriculture and the Undergraduate: Proceedings (1992), 296 pp., ISBN 0-309-04682-3 Water Transfers in the West: Efficiency, Equity, and the Environment (1992), 320 pp., ISBN 0-309-04528-2 Managing Global Genetic Resources: Forest Trees (1991), 244 pp., ISBN 0-309-04034-5 Managing Global Genetic Resources: The U.S. National Plant Germplasm System (1991), 198 pp., ISBN 0-309-04390-5 Sustainable Agriculture Research and Education in the Field: A Proceedings (1991), 448 pp., ISBN 0-309-04578-9 Toward Sustainability: A Plan for Collaborative Research on Agriculture and Natural Resource Management (1991), 164 pp., ISBN 0-309-04540-1 Investing in Research: A Proposal to Strengthen the Agricultural, Food, and Environmental System (1989), 156 pp., ISBN 0-309-04127-9 Alternative Agriculture (1989), 464 pp., ISBN 0-309-03985-1 Understanding Agriculture: New Directions for Education (1988), 80 pp., ISBN 0-309-03936-3 Designing Foods: Animal Product Options in the Marketplace (1988), 394 pp., ISBN 0-309-03798-0; ISBN 0-309-03795-6 (pbk) Agricultural Biotechnology: Strategies for National Competitiveness (1987), 224 pp., ISBN 0-309-03745-X Regulating Pesticides in Food: The Delaney Paradox (1987), 288 pp., ISBN 0-309-03746-8 Pesticide Resistance: Strategies and Tactics for Management (1986), 480 pp., ISBN 0-309-03627-5 Pesticides and Groundwater Quality: Issues and Problems in Four States (1986), 136 pp., ISBN 0-309-03676-3 Soil Conservation: Assessing the National Resources Inventory, Volume 1 (1986), 134 pp., ISBN 0-309-03649-9; Volume 2 (1986), 314 pp., ISBN 0-309-03675-5 New Directions for Biosciences Research in Agriculture: High-Reward Opportunities (1985), 122 pp., ISBN 0-309-03542-2 Genetic Engineering of Plants: Agricultural Research Opportunities and Policy Concerns (1984), 96 pp., ISBN 0-309-03434-5

Nutrient Requirements of Domestic Animals Series and Related Titles The Role of Chromium in Animal Nutrition (1997), 80 pp., 0-309-06354-X Nutrient Requirements of Beef Cattle, Seventh Revised Edition (1996), 241 pp., ISBN 0-309-05426-5; disk included Building a North American Feed Information System (1995), 70 pp. Nutrient Requirements of Laboratory Animals, Fourth Rev. Ed. (1995)., 176 pp., ISBN 0-309-05126-6 Metabolic Modifiers: Effects on the Nutrient Requirements of Food-Producing Animals (1994), 81 pp., ISBN 04997-0 Nutrient Requirements of Poultry, Ninth Revised Edition (1994), ISBN 0-309-04892-3 Nutrient Requirements of Fish (1993), 108 pp., ISBN 0-309-04891-5 Nutrient Requirements of Horses, Fifth Revised Edition (1989), 128 pp., ISBN 0-309-03989-4; diskette included Nutrient Requirements of Dairy Cattle, Sixth Revised Edition, Update 1989 (1989), 168 pp., ISBN 0-309-03826-X; diskette included Vitamin Tolerance of Animals (1987), 105 pp., ISBN 0-309-03728-X Predicting Feed Intake of Food-Producing Animals (1986), 95 pp., ISBN 0-309-03695-X Nutrient Requirements of Cats, Revised Edition (1986), 87 pp., ISBN 0-309-03682-8 Nutrient Requirements of Dogs, Revised Edition (1985), 79 pp., ISBN 0-309-03496-5 Nutrient Requirements of Sheep, Sixth Revised Edition (1985), 106 pp., ISBN 0-309-03596-1 Further information, additional titles (prior to 1984), and prices are available from the National Academy Press, 2101 Constitution Avenue, NW, Washington, DC 20418, 202-334-3313 (information only); 800-624-6242 (orders only); 202-334-2451 (fax).

Copyright © National Academy of Sciences. All rights reserved.