You are on page 1of 183

2007

Proceedings
of the

5th MID-ATLANTIC NUTRITION CONFERENCE


March 28-29, 2007 Timonium, Maryland

University of Maryland Pennsylvania State University University of Pennsylvania Veterinary School University of Delaware Virginia Polytechnic Institute and State University Rutgers, The State University of New Jersey West Virginia University & Maryland Feed Industry Council, Inc. & American Feed Industry Association, Inc. & United States Department of Agriculture
Cooperating

CONFERENCE SPONSORS
The Maryland Feed Industry Council and the Conference Organizing Committee extend sincere appreciation to the following sponsors for the generous support of this educational program for the animal feeding industry:

PATRONS Novus International


Degussa Corporation SUSTAINING SPONSOR Alltech, Inc. SPECIAL SPONSORS
Danisco Animal Nutrition Diamond V DSM Nutritional Products, Inc. Perdue Farms Incorporated Renaissance Nutrition, Inc. Southern States Cooperative Star Labs/Forage Research, Inc. The Old Mill Troy, Inc. Zinpro Corp.

SPONSORS
Alpharma, Inc. Blue Seal Feeds, Inc. C. F. Zeiler & Company, Inc. Eastern Minerals, Inc. Venture Milling Wengers Feed Mill, Inc.

2007 Mid-Atlantic Nutrition Conference


Conference Committee: Conference Chair and Editor .....................................Nick Zimmermann Program Chair............................................................Paul Patterson Symposium Chair.......................................................Curtis Novak Coordinator/Registration............................................Clare Capotosto Educational Institutions in Delaware, Maryland, New Jersey, Pennsylvania, Virginia, and West Virginia, the USDA, and the Maryland Feed Industry Council are collaborators in the planning and funding of The Mid-Atlantic Nutrition Conference.

Program Committees/Manuscript Reviewers:


General Program Committee Bob Elkin, Chair Pennsylvania State University Kathleen Crandell, Co-chair Virginia Tech Poultry Nutrition Committee Bill Saylor, Chair University of Delaware Ted Miller, Co-chair Mountaire Farms of Delaware, Inc. Roselina Angel University of Maryland Curtis Novak Virginia Tech Dairy Nutrition Committee Rich Erdman, Chair University of Maryland Clay Zimmerman, Co-chair Blue Seal Feeds, Inc. Ben Corl Virginia Tech Jim Ferguson University of Pennsylvania Tonya Gressley University of Delaware Lisa Holden Pennsylvania State University Dave Kirk Pennfield Corporation Equine Nutrition Committee Ann Swinker, Chair Pennsylvania State University Karen Engel, Co-chair Southern States Cooperative Marty Adams Southern States Cooperative Amy Burk University of Maryland Kathleen Crandell Virginia Tech Dave Marshall University of Delaware Erin Petersen University of Maryland Burt Staniar Virginia Tech Carey Williams Rutgers University

2007 Proceedings of the Fifth Mid-Atlantic Nutrition Conference


Edited by Nick Zimmermann Published in March 2007 By the Maryland Feed Industry Council Department of Animal and Avian Sciences University of Maryland College Park, MD 20742

2007 Maryland Feed Industry Council


Kathleen Crandell, President Virginia Tech Karen Engel, Vice-President Southern States Coop. Michael Elliot, Secretary Wengers Feed Mill, Inc. Mark LaVorgna Alpharma Animal Health Ted Miller Mountaire Farms of Delaware, Inc. Tim Snyder Renaissance Nutrition, Inc. Clay Zimmerman Blue Seal Feeds, Inc. Randy White Venture Milling Bob Buresh, Tyson Foods Nick Zimmermann, Treasurer University of Maryland
Conference Note: The Mid-Atlantic Nutrition Conference is a regional meeting that evolved from the Maryland Nutrition Conference for Feed Manufactures. Program content and format remain the same. ii

TABLE OF CONTENTS
Presenter Dr. Martin W.A. Verstegen Title Page The Challenges in Animal Nutrition in the 21st Century..........................................................................1 Perspectives on Feeding Athletic Horses in the 21st Century........................................................................17 Biofuels and Broilers Competitors or Cooperators?...................25 Biotechnology in the Barnyard What Will it Look Like in 2050?..............................................................................35

Dr. Raymond J. Geor

Dr. Park W. Waldroup Dr. Terry D. Etherton

Dr. Penny M. Kris-Etherton The Role of Animal Products in the Diet to Reduce the Risk of Chronic Disease: A Futuristic Vision of Potential New Foods .........................................................................41 Dr. Julia Dibner What Do We Know After 30 Years of Early Nutrition Research? ...........................................................................52 Feeding the Hen for Offspring Productivity ..................................62 The Influence of Early Nutrition: When is Day 1?........................70 Feed Manufacturing Considerations for Using DDGS in Poultry and Livestock Diets...............................................77 Process and Engineering Effects on DDGS Products Present and Future..............................................................82 Formulating Poultry Diets With DDGS How Far Can We Go? ..............................................................................91 Use of Ethanol Distillers Byproducts in Lactating Dairy Cow Diets.........................................................................100 Protein and Fertility .....................................................................109 Using Mineral and Vitamin Supplements to Enhance Reproductive Performance of Dairy Cattle......................115 Using Fats and Fatty Acids to Enhance Reproductive Performance .....................................................................116

Dr. Michael Kidd Dr. Michael Lilburn Dr. Keith Behnke

Dr. Vijay Singh Dr. Sally Noll

Dr. David Schingoethe

Dr. James D. Ferguson Dr. James N. Spain

Dr. William Thatcher

iii

Dr. Robert C. Fry Dr. Charles C. Stallings

Nutritional Management of the Organic Dairy............................130 National Resource and Conservation Service Dairy Nutrition Nutrient Management Research Projects in the Mid-Atlantic Region ..............................................140 Challenges in Determining Energy Requirements of Horses ..........................................................................147 Equine Carbohydrate Nutrition: Implications for Feeding Management and Disease Avoidance ..............................154 Protein and Amino Acid Nutrition in the Horse: Improvements and Goals for the Future ..........................162 Using the NRC to Manage Horse Nutrition.................................171

Dr. Laurie M. Lawrence

Dr. Raymond J. Geor

Dr. Patricia Graham-Thiers

Dr. David W. Freeman

All speakers were asked to submit a written paper for these proceedings. We thank the program participants for their cooperation in providing the material in this document. Manuscripts received in a timely manner were reviewed and edited by the General Program, Equine Nutrition, Poultry Nutrition, and Dairy Nutrition committees. Others were formatted for style and printed as they were received.

iv

THE CHALLENGES IN ANIMAL NUTRITION IN THE 21ST CENTURY


Martin W.A Verstegen Wageningen University Animal Nutrition Group, Marijkeweg 40 NL-6709 PG Wageningen, The Netherlands Phone: +31 317 484082 Fax: +31 317 484260 Email: martin.verstegen@wur.nl Summary Animal nutrition is increasingly an integral part of animal production. Solutions for problems in animal production can be at least partly achieved by developments in animal nutrition. In the past, this was done on one hand by determining requirements and on the other hand by determining composition of feedstuffs and diets. In recent years, mechanistic modelling is used increasingly. In addition, nutrition is used increasingly to determine biological reactions of animals. Nutrition can also become important for finding solutions for welfare and behaviour, and for environmental pollution. Of course, nutritional improvements will be used for further improvements in production efficiency. Studying the other effects of nutrition on the animal will lead to new developments in nutrients for gut health and also in nutrient genomics. In addition, developments in the field of enzymes and technology will further enable the use of fibre rich feedstuffs more efficiently. This may provide an increase in animal production by improvements achieved with the use of non-starch polysaccharides in by-products for monogastrics. Areas of the world which can not be used for the production of cereals may be used for increased ruminant production. Also refer to: Verstegen and Tamminga, 2005. Introduction Animal nutrition is an integral part of animal production. It has changed drastically as a consequence of developments in the other disciplines of animal science and also because of changes in animal husbandry practices. Developments in disciplines have been so strongly related to one another that one discipline could not have developed in isolation without developments in the other. For instance, animals have been bred to have an increased production, but the expression of this enhanced genetic potential was only possible by continuous adjustment of nutrition to the genotype of the animal. On the other hand, developments in animal properties have led to research on possible changes in nutritional needs with regard to these developments. Similar connections can be made between nutrition and housing and between nutrition and developments in preventive medicine. Many changes have taken place in animal production systems in many countries after the 1950s. As an example, changes in poultry in two decades in the Netherlands are presented in Table 1. Developments in housing have led to the general use of strawless systems in poultry, pigs and dairy cattle. More recently, however, straw is once again commonly used for some forms of production in cows and pigs. In some countries, cattle are increasingly housed in feedlot systems. In addition, developments in the nutritional sciences and related technologies have revealed insight in nutritional values and also in properties of many new agricultural products and by-products of the food industry. As an example, changes in typical diet composition of domesticated animals in the Netherlands in 1950 and 2005 are given in Table 2 for cattle and in Tables 3 and 4 for 1950 and 1988 and pigs and poultry (Vos, 1988).

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

In addition, society has become aware of consequences of animal production on the environment. In the European Union or EU, (as a follow up on initiatives of several member states) legislation is in place or in preparation on the connection of animals to surface area with regards to phosphorus (P) and nitrogen (N) excretion. The basic principle of this initiative is that no more N and P may be added than can be taken out by the plants and in addition some run-off. In addition, ammonia emissions need to be reduced drastically. So in the EU the N and P of excreta application on the land has to be in agreement with these regulations. Table 1. Changes in poultry production in the Netherlands in the two last decades of 2000. Feed efficiency is feed/gain for broilers and feed/egg mass for layers. Broilers 1980 1990 2000 Layers 1980 1990 2000 Weight (g) 1700 2000 2300 Feed (kg) 3.2 3.4 3.5 Feed/efficiency 1.88 1.70 1.50

2.50 2.20 2.05

Table 2. Rations for a dairy cow (500 kg) producing 20 kg milk with 3.75% fat (kg. day-1) in 1950 (Vos, 1988 and Dijkstra, 2005 personal communication). 1950 2005 33 % grass silage 10 kg grass hay (good quality) (29%) 17 % fresh grass 20 kg grass silage (not wilted) (58%) 25 % maize silage 2 kg dried grass (5.8%) 4 % wet byproducts 1 kg dried sugarbeet pulp (2.9%) 26 % compound feed 1.25 kg compound feed (3.65%) From Table 2 it can be derived that in this time period maize silage has replaced part of the grain. Table 3. Compound feed composition (%) for fattening pigs, 50-100 kg live weight. 1950 1988 20 17.5 30 10 4 4 6 5 1.5 maize barley rye sorghum grass meal coconut expeller soybean meal, solvent extract meat meal tankage1 Mineral and vitamin mix 13.3 37.9 3 0.2 3 10.9 1.4 15 1.7 7.5 2.2 2.5 1.4 peas tapioca meal alfalfa meal coconut expeller rapeseed meal, solvent. extract soybean meal, solvent extract hominy feed wheat middlings sugarbeet pulp cane molasses meat meal tankage1 feed fat1 Mineral and vitamin mix

It should be noted that products from the rendering are no longer allowed because of the BSE cases.

Table 4. Compound feed composition (%) for laying hens. 1963 40 13 10 4.5 5 2 5 5 2.5 2.5 3 7.5
1

1988 maize sorghum oats soybean meal, solvent extract Sunflower meal, solvent extract sesame expeller maize gluten feed wheat middlings bran alfalfa meal fish meal Mineral and vitamin mix 35 10.3 10 1.2 8.4 8.3 3 3.6 2 5 4 9.2 maize tapioca peas soybeans, heat treated soybean meal, solvent extract wheat middlings sugar cane molasses alfalfa meal feather meal, hydrolized1 meat meal tankage1 feed fat1 mineral and vitamin mix

It should be noted that these products are not allowed in the EU because of the BSE cases.

In some countries such developments have led to completely different diets to meet the changed nutrient needs of animals. In recent years, scientific progress in molecular biology has added new possibilities for biologically active substances to be used in animal nutrition. With regard to future developments in nutrition, society will increasingly put constraints on methods of use and keeping of animals and products to be used in diets for animal production. An example being put into practice is the ban on the use of all antibiotics as growth promoters in animal nutrition from January 2006 onwards by the European Union. Also in other countries developments in that direction are taking place, so alternatives need to be found. For nutrition, this means developments which involve the use of specific components of plants for the health of animals or for the quality of production. In the following, a few aspects of developments in nutrition will be pointed out and subsequently certain challenges for the 21st century will be identified and discussed. Animal nutrition will face a number of changes. It is unclear, however, in which direction animal production will develop. There are the recent developments with regard to the expanding EU and with changes as a result of developments in GATT (General Agreement on Tariffs and Trade). It will depend on how animal nutrition copes with consequences of these developments. In addition, the rapid growth in China and India will give rise to increased demand of animal products. Technical developments in nutrition itself will probably continue to focus on: a) The development of mechanistic models as new feed evaluation systems are based on available nutrients rather than energy and protein and how to incorporate the increasing knowledge on the reaction of animals to nutrition in such mechanistic models. b) The increase in feed needed for companion animals. These animals each have specific needs which will be studied more in future. c) The development and application of nutrition systems, which do not compromise animal welfare and ethics. d) The identification and investigation of components in the feed, which have a biological activity and can be used for specific functions. e) Studies on kinetics of the gastro-intestinal tract as a digestive system as well as a defence barrier to the outside world.

f) The investigation of possible functions of dietary carbohydrate fractions with regard to their influence on microflora activity and how to deal with interactions between components. g) The improvement of the prediction of digestion by further developing in-vitro techniques. h) Developments in the field of molecular biology which are related to nutrition like the development of nutraceuticals and other compounds by using techniques like nutrigenomics and proteomics. i) Nutrition and environment. How nutrition influences composition of excreta and its effects on the environment. Developments in animal nutrition until 1960 The understanding of nutrition was developing rapidly in these years. The role of nutritionists until then had been to formulate diets, rations or supplements from the knowledge of animal needs and feeding value. Around the turn of the 19th century, it was soon recognised that animals required proteins, fats and carbohydrates. Most progress until 1920 was made with regard to these nutrients as well as on energy utilisation, and less on minerals and vitamins (Church and Pond, 1982). After 1920 there was a rapid development in our knowledge in the field of vitamins, amino acids, essential fatty acids, macro and micro minerals, and energy metabolism. For nutritionists, the first priority was to prevent nutrient deficiencies. There was rapid development in the field of requirement estimates and many organisations aimed at getting the research information into feeding tables, which were then rapidly applied to feed formulation. The need for specific elements, other components for several classes and categories of animals were recognised. These tables have been often updated and provided a wealth of information for students, nutrition specialists and for the practice of animal production. In energy metabolism research was aimed at systems, which could be used to express energy value of feeds on the one hand, and energy requirements of animals at the other. These systems were intensively discussed in a whole series of symposia on energy metabolism of farm animals starting in 1958 in Copenhagen and held every three years until now. In protein metabolism, such symposia were also held in which developments in the field of protein and amino acid metabolism were discussed, first by animal species but in recent years more and more by functions like growth and reproduction. Comparative aspects between species have been focused on in recent symposia. Discussions during these symposia by researchers highlighted the need to obtain more basic and mechanistic aspects of the utilisation of energy and nutrients. In this period renewed interest in mineral metabolism led to various new research programs. These were also conducted in humans with regard to the interaction of minerals and other dietary components. Every few years the National Research Council (NRC) in the United States and Canada, and the Agricultural and Food Research Council (AFRC) in the UK and similar organisations in many other countries published revised editions of books on nutrient requirements of farm animals and other species. These were then used to determine energy and nutrient needs of animals and to derive proper diets. In the last two decades, these tables have been extended to more classes of animals, distinguishing between genotypes and different production levels. On the other hand it should be noted that until the 1970s, most studies leading to these tables were still based on the concept of considering the animal as a black box. It is now becoming more and more apparent that an animal is a complex system in which a number of subsystems are integrated and also that nutrient requirements may differ between sub-systems. For instance, in a ruminant the requirements of the rumen microbes differ widely from that of the mammary gland. Likewise, in poultry the requirement of the ovary and associated organs may be completely different from that of other organs. In all animal species the digestive tract and its requirements receives much attention,

not only because of its digestive function, but also because it is recognised to be an extremely important defence barrier to the outside world. Present developments Modelling Around 1970, scientists started to build simulation models to derive at and predict the animals responses. This development of models was done to enable prediction of growth from properties of the dietary components and of the animal. The original models were mostly deterministic and the regression approach was used most (see Moughan, 1995). However, mechanistic models also started to emerge. Nowadays, most modelling by researchers is done on the basis of the mechanistic approach. In such an approach it can be speculated that this will make most feed tables obsolete in the future (Black, 1995). However, some tables will probably continue to exist to enable students to get acquainted with quantities but also to provide the modellers with information from the tables. Also nutritionists can derive data on amounts in order to avoid non-realistic and non-possible values. The models will be based on different nutrients, which are derived from digestion in the gastrointestinal (GI) tract. Furthermore, the use of nutrients on intermediate processes for different purposes can be modelled. It should be noted that establishing nutrient requirements was mostly based on long-term studies. This meant that mostly adaptation to a diet was allowed before the measurements started. Adaptation to a diet in pigs was often one week or more and in ruminants a couple of weeks. On the other hand, on actual farms changes in diet may occur within a few days (eg. new grazing pasture) or abruptly with a new diet formulation or when a new batch of feedstuff becomes available. So in addition to the measurements after adaptation also short-term adaptation to feeds and strategies will need to receive further attention. Also short-term adaptation to climatic and other environmental conditions have hardly been studied with regard to nutrition. Until now most studies on metabolism with regard to nutrition give data on a 24 hour basis while metabolic rate will vary at least 50% within a day depending on the time scale used. Similarly, protein metabolism data on a 24 hour basis are used while the post prandial and post absorptive fate of amino acids may be different (Nolles personal communication). The rate of digestion in the GI tract determines the rate of absorption thus differences in absorption may occur as a result of asynchrony. It can be expected that the rate of hydrolyses of components in a diet can be important for variation of metabolism within a day and this may even affect efficiency of production and health of the animal. The development of mechanistic models will enable researchers to evaluate the importance of asynchrony with regard to uptake and metabolism of absorbed nutrients. Instead of digestible, metabolizable and net energy, the nutrients supplied to tissues and organs will form the basis of future systems. Development of rapid in-vitro evaluation system As a consequence of the developments in plant breeding and technology in future, feedstuffs derived there from will show much larger variation between feedstuffs with the same name compared to present-day feedstuffs. The composition, digestibility and availability of these new feedstuffs have to be known to properly evaluate them and include them into diets. New developments will probably lead to the inclusion of availability rather than digestibility. Therefore, there will be an increasing need for in rapid in-vitro test to determine these properties. Currently most of these so-called rapid bioassays are still time consuming and in-vitro or proximate-analyses based systems take at least 48 hours to complete (Leeson et al., 2000). Near infrared reflectance spectroscopy (NIRS) is a technique that determines some of the beforementioned properties and performs them within minutes. Further, according to these authors, NIRS has

the potential to be a reliable method for predicting amino-acid content and availability. This technique has been used also to determine digestible energy (DE) content in diets (see Leeson et al., 2000). Other side effects of feeding animals Issues pertaining to the environment have been at the forefront for the last few decades. In the Netherlands it started with a focus on nitrogen (N) and phosphorus (P). As a result, P emission has been reduced drastically, as shown in Table 5. Table 5. Developments in intake and excretion of N and P in pigs in kg/pig. (25-110 kg) (Jongbloed, 1999) Year In feed Feed/gain Emission P 1973 1983 1992 1996 7.4 6.2 5.0 4.7 N 23.8 24.4 26.9 26.7 3.37 3.08 2.86 2.74 P 1.62 1.18 0.77 0.67 N 4.74 4.30 4.46 4.13

This table clearly shows the enormous reduction in P emission per animal. The European Commission also focuses on regulation of N and P emission from animal production to the environment. Ammonia and N emission of all farm animals will be regulated to avoid enrichment of ground water with nitrate and P is used also to avoid eutrification. Future developments Human population increase Animal nutrition will depend on the availability of feedstuffs to provide feed to them in addition to the need of the growing human population. It is expected that the demand for meat and other animal products will sharply increase as a result of an increasing human population (See Fig 1, Pinstrup-Anderson et al., 1997). The rise in the demand for cereals and other vegetable products will however be much more than for animal products, as can be seen in Pinstrup-Anderson (1999) and Scott et al. (2000). According to these studies China will have the largest increase in demand for both meat as well as cereal products. Choct (2005) stated that in China in 1961 the average Chinese person ate 3.8 kg of meat per year. In 2004, the average is up to 54.9 kg per person. The extra animal feed for this increase has to be produced in addition to the extra need for human consumption. In India the consumption of meat is now 5.5 kg per person. The projection according to Choct is that in 2050 this will be 30.4 kg and that this will require almost 100 million more tonnes of feed. His suggestion is that now we are feeding about 70% cereal grains and 25% protein sources which is only digested by 75-80%. Why so much waste? According to him that is because Figure 1. Increase in world population, 1995-2020. S. Asia = South Asia: SSA = Sub-Saharan Africa; E & SE Asia = East and South East Asia; WANA = West Asia and North Asia; LAC = Latin America.

digesting fibre by the animal is very difficult. By using enzymes that can help digest about 500 million tonnes of fibrous by-products better, meat production could be elevated. In this respect it can be a big factor. But will the change be so big? In addition, with increasing prosperity of countries, companion animals will increase in numbers and the field of animal nutrition will increasingly study these animals. The studies of the International Food Policy Research Institute (IFPRI) mentioned above predict that it will be an increase in crop yield rather than an increase in area used for cereal production that will be responsible for the increase in total production. Combined with some of the effects suggested by Choct about increased use of enzymes to digest non-starch polysaccharides (NSP) from cereals, this can give some room for increase in production of animal products.

Figure 2. Share of increase in global demand for cereals, 1995-2020.

Figure 3. Total demand for cereals and meat products, 1995-2020.

Concomitantly with these developments, there will be an increasing pressure on animal production systems to use fewer products that can be used by humans. This means that there will be an increased pressure and need to use products for animals that are not used by humans. Requirements for food (humans) and feed (domesticated animals) are best compared and illustrated by using a common denominator (i.e. biomass maintenance and biomass production) (Table 6, Tamminga et al., 1999). Table 6. Size and annual formation of new biomass of the 1997 world population of humans and domesticated animals (calculated from FAO, 1998). Number (x109) Large ruminants Small ruminants Monogastric Fowl Total animal population Human population 1.41 1.57 1.36 13.9 18.2 5.7 Biomass (x106 tons) 332 36 47 12 427 226 Annual Production (x106 tons) 52.6 9.9 87.2 58.1 207.9 22.5

When expressed in biomass, the human population accounts for about 226 million tons. The animal population is about twice that size (427 million tons), of which nearly 85% are ruminants, which are less competitive to humans than non-ruminants. However, the estimated annual formation of new

biomass by monogastric farm animals exceeds that of the human population by a factor 7. Requirements for maintenance and biomass formation in terms of concentrates can be estimated at 1000 and 100 million tonnes for humans and 500 and 500 million tonnes for domesticated animals (Tamminga et al., 1999). Increasing animal production through pigs and poultry should therefore primarily be achieved by increasing productivity. Increasingly, and most notably in developed countries, society gives signals to animal producers that it will demand that animal production in future be animal friendly and sustainable. The human food chain, including animal products has become an ethical issue. Also the intrinsic value of animals should be recognised. Food from animals should be safe, environmentally friendly and produced in a (animal) welfare friendly system. This means that emphasis will be less on maximising production but increasingly on optimising production. It will be a matter of debate as to what is exactly meant by optimised production. Probably, this will be at such a level that marginal efficiency is still considerable, but no longer approaching zero, as is often the case nowadays. Society is signalling to producers that animal production is not just a commodity. Intrinsic values of animals should be recognised more than has been done in various intensive systems developed over the last 25 years. Also animal production will be held responsible for negative side effects of production such as the extensive loss of nutrients considered to be damaging the environment. Examples are the loss of phosphorus, nitrate, ammonia and other gases considered harmful to the environment, like nitrous oxides and methane. New techniques in animal production and animal nutrition as part of it will be able to take away the environmental side effects of animal production. However this has to be without causing other new negative side effects. Technical development, however, does not always take into account the welfare and other ethical issues that have been raised in the last few years. Challenges in future animal nutrition In addition to challenges derived from ethics, quality, health, and environment, future animal production will face many other developments. These developments derive from the earlier mentioned changes in the number of humans, amounts and types of foods for humans their food residues, and humans preferences for foods of animal or vegetable origin. Feed Properties In modern feed requirement tables often a distinction is made between different types of animals. The development of animals with specific properties, by selection or through biotechnological techniques, which enables them to produce much more meat or milk or have other properties, will require even more different feed requirement tables, if models would not be available. Knowledge on functioning of these animals and their physiology will enable to develop proper feeding practices. In combination with the study on animals which produce, for example, more protein in their body compared to lipid it is important to study which components in the natural diet have effects on partitioning of protein and lipids. Research in food physics and food chemistry in recent years has shown that many as yet unknown feed properties can influence the animal. At the start of the previous century, many laboratories determined only a few classes of components in the diet based on the Weende analyses, (i.e. dry matter, crude protein, crude lipid, crude fibre and remainder of carbohydrates as N-free extractives and crude ash). From that, determining digestibility of these components derived the potential for feed for animal products. Many feeding systems were and still are composed on that basis. Developments since then have firstly recognised the nutritional role of amino acids as components of (crude) protein. It was also recognised that in crude fat polyunsaturated fatty acids (PUFA) play a completely different nutritional

role than the saturated fatty acids. More recently it has been recognised that the carbohydrate fraction can also be divided into a range of components with different nutritional functional properties. In ruminant nutrition it has long been recognised that an important component of carbohydrates, structural carbohydrates, are essential for proper functioning of the rumen and its microbial fermentation. Further developments have been to characterize and define the quality of structural carbohydrates as Neutral Detergent Fibre (NDF), Acid Detergent Fibre (ADF) and Acid Detergent Lignin (ADL). Nowadays, non-structural carbohydrates are receiving more attention. They are divided and characterised in terms of soluble sugars and starches of different origin and differing in rumen degradation rates and hence also in rumen resistance. It was also realised that the microbial population in the foregut may have requirements for nutrients and energy different from that of the whole animal. A somewhat comparable development is foreseen for non-ruminants. In non-ruminants it was not until starch was determined routinely that interests in carbohydrates began. The determination of starch made it possible to divide the residual fraction in the Weende analysis called N-free extractives or Nfe between starch and a non-starch residue. The nutritional significance of physical and chemical aspects of this residue is now extensively explored. In connection with this, properties of feed not directly related to nutrients and nutrient density, are and will become even more important. Some of these properties are connected to physical structure, influence of feed components on activity (eg. animal welfare), influence of feed components on the activity of microorganisms in the GI tract, biologically active components in the feed, additivity of properties of different feed components in one diet, use of enzymes and immunomodulation compounds. Physical structure, has always been a component of the nutrition of ruminants and nowadays also in other animals like poultry (Bedford and Schulze, 1998). The viscosity, one aspect of physical structure, of feed components in poultry feed components are now routinely determined, since it is well known that high viscosity in a chicken diet reduces digestibility, especially in young chickens (Table 7). Similar effects of viscosity were found by Langhout (1998). Physical structure in terms of rigidity in poultry diets is also important with regard to functional properties (i.e. fineness) and also with respect to gut health. Table 7. NSP and fecal nutrient digestibility in broilers. HV* barley HV* barley & glucanase
*

Protein 77.1b 83.7a

Fat 72.3a 86.2b

Starch 95.8b 97.9

High viscosity barley (3.8% betaglucose, 13.3 mPA5 water extract viscosity) a,b values in column; (P<0.05 Almirall et al., 1995) Although in diets for pigs viscosity is of less importance than in chickens, probably due to increased water intake and increased retention time, physical properties are also of major importance. Like in poultry, fineness and particle size are important with regard to the risk of developing stomach ulcers. Processing in general will alter accessibility of carbohydrates for enzymes in the GI tract. Influence of specific components in the diet on activity and on animal welfare In animal feed evaluation systems it is generally assumed that maintenance is not influenced by the diet. Recent studies however show that inclusion of fermentable carbohydrates in the diet of pigs can change their maintenance requirements (Schrama et al., 1998, Wenk, 2001). In humans, studies have shown that food and specific components of food can affect mood and alertness. This is probably also true

for animals. It is not established yet which components in the diet or which fermentation end products are responsible for that. Future developments in research will probably show that many more components in a diet serve a specific purpose: physically, by determining conditions in the GI tract and by influencing the time after a meal at which various nutrients become available for absorption; and chemically, by influencing how much of the nutrients are digested and adsorbed. Another dietary component, which has received and is still receiving a lot of interest are the nonstarch polysaccharides (NSP). The NSP fraction of a diet contains all the carbohydrates that cannot be digested by the animal's own enzymes, but which are degraded by the microflora in the gut. In research there is increasing interest in NSP in the diet not only as NSP itself but also about the possibilities to use each component of NSP in human and animal diets. The interest in the role of dietary NSP is increasing because of the high priority and demand of the cereals for human consumption. Increased use of cereals for human consumption will result in more by-products becoming available. These by-products often have a high level of NSP. Until now, one energy value has been given for fermentable non-starch polysaccharides in diets for pigs. From different studies, it is questionable whether one value for the utilization of different digestible non-starch polysaccharides for pigs can be used. Recently Rijnen (unpublished results) showed that fermentable NSP of different origin (NSP from sugarbeet pulp, soybean hulls and coconut meal) each may have a different effect on physical activity of pigs. Specific effects of feed components on gut microbiology Another aspect of NSP is its connection to microbial activity. Dietary manipulation of microbial activity by specific carbohydrates is a topic which will get much more attention in the coming years. Especially with the ban on Antibiotics as Growth Promoters (AGP) imposed in the European Union. The interest in other methods to manipulate microbial life in the GI tract has increased and will increase further. Anderson et al. (1999) suggested that indigenous micro biota in the small intestine can depress the amount of nutrients to be absorbed from the diet. One aspect of this is that the micro flora will compete with the animal for nutrients and secondly that toxic metabolites will affect the gut wall and increase its turnover. Both of these aspects could potentially contribute to reduced growth. For example, Vervaeke et al. (1979) suggested that 6% of the net energy in a pigs diet could be lost due to bacterial utilization of glucose in the small intestine. It has also been shown that blocking urease activity in the GIT, which reduces ammonia release from urea, increased growth. Amidst the uncertainty, there is at least sufficient evidence to show that the growth promoting effect of antibiotics is due to a mechanism in the GIT and not elsewhere in the body. Antibiotic growth promoters are thought to reverse this microbial-induced growth depression. This means that there is an increased availability of nutrients and/or reduction in the maintenance costs of the gastrointestinal system. In the search for alternatives to antibiotics, it would therefore be logical that such alternatives would also act according to one of these two mechanisms. Until now, our knowledge of how we can stimulate beneficial microflora in the GI tract is very limited (Williams et al., 2001). It can be expected that some carbohydrates are the primary sources of prebiotics as alternatives to antibiotics. These carbohydrate compounds can be considered as potential alternatives for AGP when they directly or indirectly favor or mimic one of the actions of AGP. Prebiotics have been defined as nondigestible food ingredients that beneficially affect the host by selectively stimulating the growth and/or activity of one or a limited number of bacteria in the colon leading to an improvement in host health

10

(Gibson and Roberfroid, 1995). The purpose of prebiotics is therefore to provide a substrate for beneficial GIT microbes (e.g. Bifidobacterium spp.; Lactobacillus spp.). Most of the compounds investigated to date, have been a form of carbohydrate. As indicated earlier, in traditional feed characterization (e.g. proximate analysis), no differentiation is made between different classes of carbohydrates. However, for cereal grains, at least 80% of the components are carbohydrates, of which 70-90% are composed of starch. Thus, the remaining fraction of which NSP is an important component can make up 10-30% of the carbohydrates present in grain. Part of these ingredients belonging to NSP is soluble and part is non-soluble. The large intestinal bacteria are probably well adapted to dynamic changes in their nutrient supply. Some are more specialized in the hydrolysis of plant polysaccharides and may produce small molecular weight carbohydrates from large polymers. Non-starch polysaccharides are increasingly being studied, both in relations to their effect on GIT function and on the volatile fatty acids, which are produced. Certain carbohydrates are now recognized as having prebiotic activity in the large intestine. It has been estimated that 40-60% of non-digestible oligosaccharides (NDO) and up to 20% of the other NSP, are actually fermented in the small intestine of pigs. This confirms that the high numbers of active bacteria present in the small intestine are potentially capable of fermenting carbohydrates. Fermentation in the small intestine has been shown to occur in piglets (Houdijk, 1998) and in young chickens (Smits, 1996). From studies in Wageningen (see Alles, Houdijk, Van Laere, and Hartemink: In Hartemink, 1997) it became clear that many NDO might be fermented too quickly. In piglets, some NDO are completely fermented before reaching the large intestine (Houdijk, 1998). Under these conditions, bacteria in the large intestine are forced to use protein as an energy source leading to an increased production of branched-chain fatty acids and ammonia. Thus future research may focus at a better control of fermentation in the GIT, with the aim of a more continuous fermentation of carbohydrates along the entire GIT. Houdijk (1998) concluded that NDO as such, have so far shown little growth promoting effect, but may still contribute in some cases, to better GIT health. The importance of location of the fermentation of the prebiotics (e.g. NDO) has been demonstrated and various approaches have been used to study this. Mostly they require animal experiments which are sometimes invasive. New in-vitro techniques will also be developed which may mimic a location of fermentation or a site of fermentation. One such method has developed by Minekus (1998) for humans and pigs and by Smeets et al. (1999) for dogs. Their model can simulate the dynamic conditions in the stomach and small intestine by mimicking; a) physiological transit time and peristaltic movements; b) concentrations of electrolytes and enzymes; c) pH values in different parts and d) absorption of digested products. They also developed a large intestinal section of the model. It was shown that inulin, soy polysaccharides and fructo-oligosaccharides (FOS) were already partly fermented at the beginning of the large intestine. Inulin and arabic gum were completely fermented mid-way along the large intestine. On the other hand, resistant starch and -cellulose were fermented only in the latter part of the large intestine. Houdijk (1998) discussed why a continuous fermentation throughout the large intestine may be beneficial. In each segment of the GIT there has to be an appropriate carbon to nitrogen (C/N) ratio (Borg-Jensen, 1993). There is considerable evidence that under non-optimal conditions of husbandry, NDO may have a beneficial effect on young animals. Also, the combination of NDO and probiotics could be beneficial under these circumstances (Mul, 1997). It can be hypothesized that a combination of various substances with different rates of fermentation will be effective in mimicking some of the antibiotic effects which have been suggested by Anderson et al. (1992).

11

If a substrate is rapidly fermentable so that most of the fermentation takes place in the small intestine, insufficient carbohydrate would then be available as an energy source for bacteria in the large intestine. This results in the fermentation of protein (other microorganisms, digestive enzymes, sloughed cells etc.) which, as was already described, will lead to the production of more Branch Chain Fatty Acids (BCFA) and release of ammonia. Indeed, it has been shown that in piglets fed easily fermentable NDO, more BCFAs and ammonia were present in the intestinal chyme. This is considered negative for the animal (Anderson et al., 1999). At our laboratory, Bauer et al. (2001) performed a large study investigating the fermentation characteristics of a range of different carbohydrates. Using a modified cumulative gas production technique (Williams et al., 1995), fermentability was assessed according to the kinetics of gas production, VFA and ammonia production, and pH of the medium at the end of fermentation. The inoculums were taken from pigs, which had not received any of the substrates to be tested. Table 8 shows the half-time (T) of maximum gas production (a measure of fermentation kinetics), and the ratio of branched to straight-chain fatty acids (BCR) of various tested carbohydrates as substrates in vitro. Table 8. T (half-time for asymptote of gas production -h) during fermentation of some products and ratio of branched to straight-chain fatty acids (BCR) (Bauer et al., 2001). Feed Ingredient T (h) BCR ab Arabic Gum 24.5 0.079bc Guar Gum 16.4b 0.082c a Xylan 29.5 0.112b b FOS 16.8 0.062c b TOS (Trans-galactooligosaccharide) 16.1 0.137b b Sugarbeet pulp 15.1 0.094c ab Jerusalem artichoke inulin 22.9 0.127b Chicory inulin 21.6ab 0.143 a,b,c Superscripts which differ in the same column are significantly different (P<0.05) The results show that there can be large differences in the fermentability of different carbohydrates, both in terms of the rate of fermentation and also in the formation of end-products. Until now results of most studies on the beneficial effects of some carbohydrates are disappointing because the carbohydrates are disappearing from the gut too quickly. Therefore the rate of fermentation seems to be crucial. It can be expected that enzymes can be made that are specific to beneficial carbohydrates with respect to site, rate of fermentation, and species of micro flora to be influenced. Additivity Currently additivity of the feeding value of different feed components in feed evaluation systems is a basic assumption. Also the values for the feed and the requirements for animals should be expressed in the same way (parameter). The assumption of additivity has made it easy for nutritionists and feed manufacturers to compose diets from a large variety of ingredients. With studies on kinetics of digestion it has become clear that additivity does not always occur. In a large variety of studies (eg. Rerat (1985), Graham et al. (1966), Bakker (1996) and Smits (1986)) it has been shown that increased levels of NSP in a diet for non-ruminants decreases digestibility and absorption of nutrients from starch, protein and fats. The mechanisms for this are not completely understood but several have been suggested by de Lange (2000): a) endogenous losses in the small

12

intestine are increased with increased NSP (secretions, enzymes, mucus and the sloughing off of gut wall cells); b) viscous NSP in particular will reduce movement of enzymes and nutrients in the digesta and the mixing of these; c) feed components with NSP will mostly have some nutrients enclosed inside the cell walls, in this way the NSP physically prevents or limits the access of enzymes to the components to be hydrolyzed; d) microbial activity is changed and some toxins are produced in the small intestine; and e) alteration of morphology of the small intestine is mentioned in various studies (Bakker 1996). Bakker (1996) showed that the ileal and fecal digestibility of some nutrients are not predicted well from pure ingredients. Also for amino acids there may be a lack of additivity of apparently absorbed amino acids when low protein feeds are mixed with high protein ingredients. Nyachoti et al. (1997) and de Lange and Fuller (2000) concluded that by correcting apparent ileal digestible amino acids properly for endogenous amino acids can eliminate that part of non-additivity. However, the impact of microbial activity in the small intestine on the estimation of absorbed amino acids, nutrient and endogenous, from small intestine is not well known. If the kinetics of digestion and product to be produced can be properly quantified then they may be incorporated into models to predict nutrients absorption and the use of these in metabolism. Environmental Aspects Until recently, nutrition has solely focused on the amounts of nutrients that can be derived from ingested feed, so nearly all new feeding tables present the amounts of nutrients in feedstuffs. They also increasingly provide estimates of nutrients apparently absorbed from a diet and also estimates of nonabsorbed nutrients that can be found. Excreta was not considered to be a problem until recently. Nowadays, in many countries or areas farmers have to focus on ways to avoid adding minerals via excreta to the environment, which has become a burden. Most research has focused on the amounts of minerals and nitrogen (N), which are excreted, and not with the quality of the excreta (Figure 4). The quality of manure will become important issues in the 21st century. Some developments already show this. It is well known that by allowing proper microbial fermentation in the large intestine, more microbial biomass will be present in the excreta. This results in apparently lower amounts of digestible N, even though ileal digestible N may be the same. The other consequence is that there is less N in the urine. This is only true if fermentation continues in the large intestine of pigs (Houdijk, 1998). As a consequence, the C/N ratio in the manure is much higher, which is considered beneficial for Figure 4. Nitrogen cycling in nature. soil life. Moreover, loss of N from N2O N2 NH3 the excreta may be less. Until now most emphasis has been on reducing N-fixation volatilisation total N in manure. It can be NH3 expected that by combining both, a reduced excretion and a shift from denitrification urinary excretion to fecal excretion, N2O HN4+ reducing N in feed will result in an increase in fermentative compound nitrification PLANT production. This will optimize the NO2quality of the manure (Canh 1998). This aspect may hold for ruminants NO WASTE ANIMAL as well as the non-ruminants.
3

SOIL, WATER

One of the biggest challenges of animal production is to have animals produce in a

13

sustainable way. This also holds true for a selection of those animals that are robust enough to stay healthy and keep producing. In recent years, some developments have taken place, which show that more specific attention can be placed on these animals. One such development was the occurrence of ascites in broiler chickens (Scheele 1996), Scheele et al. (1999). Studies clearly showed that in broiler chickens ascites developed as a result of impaired oxygen transport capacity associated with low feed/ gain ratio. Limiting the lysine in the diet clearly reduced ascites in one week in a sensitive strain of chickens. No effect was seen in a non-sensitive strain. Scheele (1996) explained this as a problem with selection against the maintenance level in chickens. This agrees with findings of Luiting (1991) who showed that in ageing hens, the animals with lower residual feed intake (low maintenance) showed much more signs of stress (losing feathers) than the animals with a high residual intake. One of the earlier studies of Vercoe and co-workers in the 1970s in Australia similarly showed that selection for high efficiency (high heat tolerance) in beef cattle was only effective when a vigorous scheme of deworming was applied. If this was not applied, animals suffered much more from the parasites than controls. This may mean that when selecting for high efficiency the selection against adaptability and/or resistance must be avoided. The animal needs sufficient metabolic scope for adaptation in addition to high production. Conclusions Developments in animal nutrition and related fields will present many opportunities to help adjust animal production to address new regulations and other challenges. Animal nutrition can contribute to: Production systems which do not compromise animal welfare and ethics Reduction in undesired behaviours Gut health Development of in vitro nutrition testing systems Developments in application of nutrigenomics Solutions for environmental solutions References Almirall, M., M. Francesh, A.M. Perezvendrell, J. Brufau, and E. Estevegarcia. 1995. The difference in intestinal viscosity produced by barley and beta-glucanase alter digestive enzyme activities and ileal nutrient digestibilities more in broiler chicken than in cocks. J. Nutr. 125:971-951. Anderson, D.B., V.J. McCracken, R.T. Aminov, J.M. Simpson, R.J. Mackie, M.W.A. Verstegen and H.R. Gaskins, 1999. Gut microbiology and the mechanism of action of growth promoting antibiotics in swine. Pigs News & Information 20:115N-119N. Bakker, G.C.M., 1996. Interaction between carbohydrates and fats in pigs. PhD Thesis, Wageningen University, The Netherlands. Bauer E., B.A. Williams, C. Voigt, R. Mosenthin, and M.W.A. Verstegen. 2001. Microbial activities of faeces from unweaned and adult pigs, in relation to selected fermentable carbohydrates Animal Science 73:313-322. Bedford, M.R. and H. Schulze, 1998. Exogenous enzymes for pigs and poultry. Nut. Res. Rev. 11:91-114. Black, J.L., 1995. Modeling Energy metabolism in the pig Critical evaluation of a simple reference model. Pages 87-102. In: P.J. Moughan, M.W.A. Verstegen and M.I. Visser-Reyneveld, eds. Modeling Growth in the Pig. Wageningen Press, The Netherlands. Borg Jensen B., 1993. The possibility of manipulating the microbial activity in the digestive tract of monogastric animals. Proceedings 44th EAAP meeting. August 16-19 Foulum, Denmark.

14

Canh, T.T., 1998. Ammonia emission from excreta of growing-finishing pigs as affected by dietary composition. Ph.D. Thesis Wageningen University, The Netherlands. Church, D.C. and W.G. Pond. 1982. Basic Animal Nutrition and Feeding. 2nd Ed., John Wiley & Sons, Brisbane. De Lange, C.F.M. and M.F. Fuller. 2000. Advances in feed evaluation for pigs. Pages 221-241. In Moughan, P.J., M.W.A.Verstegen, M.I.Visser-Reyneveld, eds., Feed Evaluation, Principles and Practice. Wageningen Press, The Netherlands. De Lange, C.F.M. 2000. Characterization of non-starch polysaccharides. Pages 77-92. In: Moughan, P.J., M.W.A.Verstegen, M.I.Visser-Reyneveld, eds., Feed Evaluation, Principles and Practice. Wageningen Press The Netherlands. Gibson, G.R. and M.B. Roberfoid, 1995. Dietary modulation of the human colonic microbiota: introducing the concept of prebiotics. J. Nutr. 125:1401-1412. Graham, H., K. Hesselman, and P. Aman. 1986. The influence of wheat bran and sugar-beet pulp on the digestibility of dietary components in a cereal-based pig diet. J. Nutr. 116:242251. Hartemink, R., 1997. (Compiler) Non-digestible oligosaccharides: healthy food for the colon. Pages 1177. Wageningen Acad. Publ., The Netherlands. Houdijk, J.G.M., 1998. Effects of non-digestible oligo saccharides in young pig diets. PhD Thesis Wageningen University, The Netherlands. Jongbloed, A. 1999. PHLO postgraduate course on developments in pig nutrition [in Dutch]. Wageningen University The Netherlands. Langhout, D.J., 1998. Role of microbiota as affected by non-starch polysaccharides in broiler chicks. PhD Thesis Wageningen University. The Netherlands. Leeson, S., E.V. Valdes and C.F.M. de Lange. 2000. Near infra-red reflectance spectroscopy and related technologies for the analysis of feed ingredient p. 93-109. In Moughan, P.J., M.W.A.Verstegen, M.I.Visser-Reyneveld, eds., Feed Evaluation, Principles and Practice. Wageningen Press, The Netherlands. Luiting, P., 1991. The value of feed consumption data for breeding in laying hens. Ph.D. Thesis Wageningen University The Netherlands. Minekus, M. 1998. Development and validation of a dynamic model of the gastrointestinal tract. Ph.D. Thesis. University of Utrecht, The Netherlands. Moughan. P.J., 1995. Modeling protein metabolism in the pig critical evaluation of a simple reference model p103-124 In: P.J. Moughan, M.W.A. Verstegen and M.I. Visser-Reyneveld, eds., Modeling Growth in the Pig. Wageningen Press, The Netherlands. Mul, A. J. 1997. Application of oligosaccharides in animal feeds. Page 106. In: Proc. Int. Symp. "Nondigestible Oligosaccharides: Healthy Food for the Colon?", R. Hartemink, ed., Wageningen Acad. Publ., The Netherlands. Nyachoti, C.M., J.L. Atkinson and J. Leeson 1991. Sorghum tannins a review. Worlds Poult. Sci. J. 53:5-21. Pinstrup-Anderson, P., R. Pandya-Lord and M.W. Rosegrand.1997. The world food situation: Recent developments emerging Issues and Long Term Prospects. Report IFPRI, Washington, D.C. 1997. Pinstrup-Anderson, P., R. Pandya-Lord and M.W. Rosegrand. 1999. Critical issues for the Early TwentyFirst Century. Report IFPRI, Washington, D.C. 1999. Rerat, A.A. 1985. Intestinal absorption of end products of digestion of carbohydrates and proteins in the pig. Archiv fr Tierernhrung 35:561-580. Scott, C.J., M.W. Rosegrand and C. Ringlei. 2000. Roots and tubers for the 21st Century. Trends, Projects and Policy Options. Report IFPRI, Washington D.C. 2000. Scheele, C.W. (1996). Ascites in chickens. Ph.D. Thesis Wageningen University. The Netherlands. Scheele, C.W., M.W.A. Verstegen, R.P. Kwakkel, R.A. Dekker and C. Kwakernaak 1999. Towards a new net energy system for poultry in The Netherlands. Proc. 12th European Symposium on Poultry Nutrition, Veldhoven, The Netherlands Branch of WPSA 399-407.

15

Schrama, J.W., M.W. Bosch, M.W.A. Verstegen, A.H.P.M., Vorselaar, J. Haaksma, and M.J.W. Heetkamp. 1998. The energetic value of nonstarch polysaccharides in relation to physical activity in group-housed, growing pigs. J. Anim. Sci. 76:4016-3023. Smeets-Peeters M.J.F., M. Minekus, R. Havenaar, G.J. Schaafsma and M.W.A. Verstegen. 1999. Description of a dynamic in vitro model of the dog gastrointestinal tract and an evaluation of various transit times for protein and calcium. Alternatives to Laboratory Animals (ATLA) 27:935-949. Smits, C.H.M., 1996. Viscosity of dietary fiber in relation to lipid digestibility on broiler chicken. PhD Thesis Wageningen University, The Netherlands. Tamminga, S., H.J.M. Udo, M.C.J. Verdegem and G. Zemmelink, 1999. The role of domesticated farm animals in the human food chain. In: Quantitative approaches in Systems Analysis, AB-DLO\PE, Wageningen 21:67-78. Theodorou, M.K., B.A. Williams, M.S. Dhanao, A.B. McAllan and J. France, 1994. A simple gas production method using a pressure transducer to determine the fermentation kinetics of ruminant feeds. Anim. Feed Sci. Tech. 18:185-197. Verstegen M.W.A and S. Tamminga, 2005. The Challenges in animal Nutrition in the 21st Century, Pages 3-30. In: New Challenges in 21st Century Animal Nutrition. Babinsky, L., ed., Proc. 12th Int. Symp. on Animal Nutrition, University of Kaposvar. Vervaeke, T.J., J.A. Decuypere, N. Dierick and H.K. Hendericks, 1979. Quantitative in vitro evaluation of the energy metabolism influence by virginiamycine and spiramycin used as growth promoters in pig nutrition. J. Anim. Sci. 49:486-496. Vos, M.P.M. 1988. The impact of energy metabolism research on feed evaluation and animal feeding in the Netherlands. Pages 20-28. In: Proceedings 11th Symposium on Energy Metabolism of Farm Animals, EAAP publication nr. 43, PUDOC, Wageningen, The Netherlands. Wenk, C. 2001. The role of dietary fiber in the digestive physiology of the pig. Anim. Feed Sci. Tech. 90:21-33. Williams B.A., A.F.B. van der Poel, H. Boer and S. Tamminga. 1995. The use of cumulative gas production to determine the effect of steam explosion on the fermentability of two substrates with different cell wall quality. J. Sci. Food Agric. 69:33-39. Williams B.A., M.W.A. Verstegen and S. Tamminga. 2001. Fermentation in the monogastric large intestine: its relation to animal health. Nut. Res. Rev. 14:207-227.

16

PERSPECTIVES ON FEEDING ATHLETIC HORSES IN THE 21ST CENTURY


Raymond J. Geor, BVSc, PhD Department of Animal & Poultry Sciences, Virginia Tech Middleburg Agricultural Research and Extension Center 5527 Sullivans Mill Road Middleburg, VA 20117 Office: 540-687-3521 x26 Fax: 540-687-5362 Email: rgeor@vt.edu Summary While nutrition is unlikely to compensate for a lack of natural ability or inadequate physical training, there is a general belief that appropriate nutritional management is necessary for any horse to achieve its athletic potential. For racehorses and other elite performers (e.g. 3-day event or endurance horses), energy needs are as much as twofold higher relative to the non-working state. This high energy requirement represents the greatest challenge in the design and application of feeding programs for athletic horses. Traditional feeding programs emphasize use of high starch grains or grain by-products that potentially increase the risk for digestive tract disorders including gastric ulcers and colic. The more modern approach utilizes alternative sources of energy such as vegetable oils and feeds rich in non-starch polysaccharides (e.g. sugar beet pulp, soya hulls). Inclusion of these alternative energy sources facilitates a reduction in the level of starch feeding without compromising the caloric density of the ration. This approach is particularly important in the management of horses with chronic exertional rhabdomyolysis. There is some evidence that fat or oil supplementation of horses in training may enhance performance during exercise requiring single or repeated, high-intensity efforts. Enhancement in the capacity for fat oxidation following fat adaptation also may be beneficial for endurance-type exercise. The goal of future research should be development of feeding practices that reduce risk of disease and injury rather than boost the performance of athletic horses. Introduction Horses are used for a wide range of athletic pursuits for which the primary objective is winning. While nutrition is unlikely to compensate for a lack of natural ability or inadequate physical training, there is a general belief that appropriate nutritional management is necessary for any horse to achieve its athletic potential. That said, what constitutes an ideal diet is far from black and white, and feeding practices tend to be more based on tradition than evidence from well-designed scientific studies. Nonetheless, in the past 15 or so years there has been a considerable body of work regarding the effects of different sources of dietary energy on metabolism and athletic performance in horses, and the results of these studies have influenced feeding practices. This paper briefly reviews the physiological basis for the horses elite athletic ability, discusses contemporary information on the health and performance effects of the different sources of dietary energy fed to athletic horses, and provides some perspective on areas for future research. The Horse as an Athlete For its size, the horse is an extraordinary athlete, a characteristic that is the result of evolution of horses as grazing animals on the ancient prairies of North America. Survival in these open lands was enhanced by speed, to escape predators, and endurance, required to travel long distances in search of feed and water. These attributes are shared by pronghorn antelopes, another species that evolved on the prairies. The equid characteristics of speed and endurance were subsequently modified or enhanced by
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

17

selective breeding by humans. Horses have been bred or adapted to a large variety of uses. Large, heavy breeds of horses were bred for draft work, such as pulling plows, sleds or carts, or military work, such as the chargers that carried heavily armored knights into the battles of the Middle Ages. Lighter horses were bred for speed and endurance and were used for transportation, herding and sport. Thoroughbred race horses run at high speed (18m/s, 64km/h) over distances of 800 to 5000 meters, Standardbred horses trot or pace at high speed for distances up to 3600m, Quarter Horses sprint for 400m or less at speed as high as 88km/h, sometimes around figure of eight courses delineated by barrels (barrel racing), and Arabians trot or canter for up to 160km in a single day during endurance events (and over longer distances during multi-day races). The athletic capacity of horses is attributable to a number of physiologic adaptations (Hinchcliff and Geor, 2004). In some cases these adaptations are not affected by training, for example lung size, whereas others change in response to training, for example blood volume and maximal aerobic capacity (VO2max). The superior athletic ability of horses is attributable to their VO2max, large intramuscular stores of energy substrates (in particular glycogen), high mitochondrial volume in muscle, the ability to increase oxygen-carrying capacity of blood at the onset of exercise through splenic contraction, efficiency of gait, and efficient thermoregulation. The VO2max of horses is approximately 2.6 times that of similarly sized cattle, and on a body weight basis is approximately 22.5 times that of highly trained men (Hinchcliff and Geor, 2004). Horses have structural adaptations that enhance oxygenation of blood in the lungs, oxygen transport capacity of blood, and the ability to deliver oxygen to tissues. The large aerobic capacity in horses is associated with a number of factors including a larger maximum cardiac output and higher hemoglobin concentration (Hinchcliff and Geor, 2004). The presence of large amounts of readily available substrate, in particular glycogen, in close proximity to mitochondria is another factor that contributes to the horses elite athletic ability. The glycogen concentration in horse muscle (homogenate of middle gluteal m.) is approximately 130-150 mmol/kg wet weight (550-650 mmol/kg dry weight), which is considerably higher when compared to sled dogs (70-80 mmol/kg ww or 330-350 mmol/k dw), or humans (80-140 mmol/kg ww) (Hinchcliff and Geor 2004). Glucose availability in skeletal muscle may be an important determinant of performance in this species during sustained exertion, in addition to the demonstrated effect of glycogen depletion during intense exercise (Lacombe et al., 2001). The rate of post exercise muscle glycogen replenishment is much slower in horses when compared to humans and other species. In humans, the ingestion of hydrolyzable carbohydrate at a rate of 0.7-1.0 g/kg bwt every 2 hours results in muscle glycogen synthetic rates of 5-8 mmol/kg muscle/h during the 6-12 hour period after glycogen-depleting exercise, and complete glycogen replenishment is achieved within 24 hours (Kiens, 2001). By comparison, when hydrolyzable carbohydrate (e.g. glucose or glucose polymers) is provided orally the maximum rate of muscle glycogen synthesis in horses is approximately 1-2 mmol/kg muscle/h, and as much as 48 to 72 hours is required for complete replenishment (Lacombe et al., 2004; Geor, 2006). As discussed below, the horse appears to have a limited capacity for the small intestinal digestion of hydrolyzable carbohydrates, and this may limit systemic glucose availability thereby restraining the rate of muscle glycogen resynthesis. Meeting the Energy Requirements of Elite Equine Athletes In general, the role of nutrition and feeding in the management of athletic horses is to satisfy nutrient requirements while maintaining health and well-being. However, actual feeding strategies may vary widely because of the different demands of training and competition. For example, whereas the consumption of a high forage diet may be of benefit to an endurance horse via promotion of a large fluid reservoir in the hindgut, this strategy is not desirable in a racehorse because excess gut fill may be energetically disadvantageous during high-intensity exercise. Similarly, the rationale for nutritional intervention immediately before and/or during competition will differ between disciplines. For an endurance athlete, feed consumption before and during a race may enhance energy supply and delay the

18

onset of fatigue. On the other hand, for a racehorse the consumption of a meal within 1-2 hours of the race start is less likely to substantially impact energy use and athletic performance. For racehorses and other elite performers (e.g. 3-day event or endurance horses), energy needs are as much as twofold higher relative to the non-working state (National Research Council, 2007). This high energy requirement represents the greatest challenge in the design and application of feeding programs for athletic horses. Traditional feeding programs have emphasized use of high starch grains or grain byproducts that potentially increase the risk for digestive tract disorders including gastric ulcers and colic. Poor appetite, particularly in racehorses, can also hamper efforts to supply the nutrients required for maintenance of body weight, condition and performance. Grain-Associated Disorders As a non-ruminant herbivore, horses evolved to utilize forages high in structural carbohydrates, with bacterial fermentation and production of volatile fatty acids in a highly developed large intestine. However, most often forage alone will not meet the energy needs of horses in hard athletic training. Traditionally, this energy deficit has been overcome via the feeding of feeds rich in starch (i.e. cereal grains and/or grain by-products), often with a concomitant decrease in the provision of forage when compared to the non-working state. For example, survey studies have indicated that racehorses weighing 450 to 550 kg typically receive 4 to 6 kg of feed per day (40-50% of dry matter intake), with some horses receiving more than 8 kg per day (70% of DM intake) (Southwood et al., 1993). This approach is at odds with the digestive physiology of the horse. The horse has limited capacity for digestion of hydrolyzable carbohydrates (particularly starch) in the small intestine. Large starch-rich meals may overwhelm the digestive capacity of the small intestine and promote the flow of undigested hydrolyzable carbohydrate to the large intestine. This not only reduces the efficiency of feed utilization, but also increases the risk for digestive disturbances associated with excessive and uncontrolled fermentation of the undigested hydrolysable carbohydrate in the large intestine. Epidemiologic studies have demonstrated that grain feeding is a risk factor for colic, a costly and potentially fatal condition of horses. For example, Tinker et al. (1997) found a 5- to 6-fold increase in the risk for a colic episode when horses received more than 2.5 kg of feed per day. Similarly, Hudson and colleagues (2001) reported that the feeding of more than 2.7 kg oats per day or a recent change in grain feeding (within the last 2 weeks) was associated with increased risk of colic. Gastric ulcer disease, a condition with high prevalence in athletic horses (80-90% of Thoroughbred racehorses in training), has also been associated with the feeding of a high grain, low forage diet (Andrews et al., 2005). There is substantial fermentative activity in the stomach of the horse after consumption of starch-rich meals, evidenced by increased lactate and VFA concentrations in gastric fluid. Furthermore, VFAs are injurious to gastric squamous mucosa particular under the prevailing acidic conditions within the stomach (pH 4.0) (Nadeau et al., 2003). The size of grain meals may affect the extent of intra-gastric fermentation and thus VFA production. Mtayer et al. (2004) compared gastric emptying rate in horses fed a small (300 g/100 bwt) vs. large (700 g/100 kg bwt) high-starch feed. Although the calculated rate of gastric emptying (g/min) was higher with the large meal, gastric emptying in terms of percent of the total meal was much slower. Thus, with large starch-rich meals intra-gastric VFA production may be favored due to the large load of fermentable substrate and longer residence time in the stomach. Two forms of chronic exertional rhabdomyolysis (polysaccharide storage myopathy [PSSM] of Quarter horses and related breeds, and recurrent exertional rhabdomyolysis [RER] of Thoroughbreds) also have been associated with high starch (grain) diets (Valberg, 2006). High starch (and sugar) diets likely contribute to the pathogenesis of PSSM by enhancement of glucose uptake and storage (as glycogen) in skeletal muscle. Thoroughbred horses with a nervous temperament have a higher incidence of rhabdomyolysis than calm horses, and it is possible that diets with high starch contribute to disease expression via exacerbation of anxiety and excitability (Valberg, 2006).

19

Recognition of the association between grain (starch) feeding and gastrointestinal and muscle diseases has focused attention on the development of feeding strategies that mitigate risk of disease, while still meeting energy requirements. First, it is advisable to limit the size of individual grain-based meals to avoid starch bypass to the large intestine (e.g. no more than 2.0 kg feed per meal for a 450-kg horse). Second, only cereal grains with high pre-cecal starch digestibility should be included in energy concentrates for horses. Whereas oat starch (at up to 3 g/kg bwt/meal) has a pre-cecal digestibility of greater than 90%, approximately 35% of an equivalent dose of cornstarch reaches the cecum undigested (Radicke et al., 1991; Meyer et al., 1993). Similarly, the pre-cecal digestibility of unprocessed barley is substantially lower when compared to oats (Cuddeford, 2001). However, heat treatments such as micronisation, extrusion and steam flaking significantly improve the pre-cecal starch digestibility of barley and corn. The preferred strategy for mitigation of problems associated with the feeding of high starch feeds is to make more use of vegetable oils and sources of non-starch polysaccharides (e.g. sugar beet pulp, soya hulls). Inclusion of these alternative energy sources facilitates a reduction in the level of starch feeding without compromising the caloric density of the ration. This approach is particularly important in the management of horses with chronic exertional rhabdomyolysis (Valberg, 2006). The ideal amount of dietary oil for athletic horses has not been determined, and is an area deserving of further research. In many of the studies that have examined the effects of dietary fat or oil on metabolic responses to exercise, the ration provided approximately 20% to 25% of DE from fat (approximately 10% fat on a total diet basis). This level of fat supplementation appears to be considerably higher than that practiced by horse owners and trainers. In one survey of feeding practices applied to endurance horses, approximately 55% of horses were fed additional fat in the form of oil or rice bran. However, on a total diet basis, the average percentage fat was 2.3%, with a range of 1.45% to 6.9% (Crandell, 2002). Nutrition and Athletic Performance Do particular diets, supplements or feeding strategies confer athletic performance advantages? Certainly, the marketplace is overflowing with special supplements that promise to provide the equine (and human) athlete with a performance advantage, either by boosting athletic capacity or by mitigating a problem that may impair performance (e.g. osteoarthritis). Furthermore, use of these types of dietary additives is widespread (Harris, 1999), perhaps reflecting the desire of owners, trainers and riders to gain a competitive edge and/or assure the general health and well-being of horses under their care. Supplements may include essential nutrients or substances purported to have a role in metabolism (e.g. creatine, carnitine, branched-chain amino acids) or tissue function but that are not recognized as an essential nutrient (e.g. chromium, coenzyme Q10, herbal products, glucosamine). As reviewed elsewhere (Harris and Harris, 2005; Geor, 2006) for most, if not all, of these supplements there is little or no scientific evidence of efficacy in horses. In this authors opinion, the key word in nutrition supplement is nutrition. In the context of feeding athletic horses, the addition of a supplement to the ration may be justifiable on nutritional grounds, but rarely on the basis of improved athletic performance. Of course, this view is unlikely to curb this highly lucrative sector of the equine feed industry. Another area of ongoing interest is the influence of dietary macronutrients (especially carbohydrates and fats) on exercise performance. This area has received considerable attention in human sports nutrition. For example, it is universally accepted that carbohydrate availability in skeletal muscle is an important determinant of exercise performance, particularly during moderate intensity exercise lasting 1 hour or more, and feeding strategies (glycogen loading) that increase pre-exercise muscle glycogen content enhance endurance performance of human athletes (Hawley et al., 1997). Carbohydrate Loading The original glycogen loading protocols, pioneered by Bergstrm and colleagues (1967), involved a 3-4 day phase of hard training and a low carbohydrate diet for induction of muscle glycogen

20

depletion, followed by a 3-4 day loading phase of high carbohydrate intake and exercise taper. This protocol resulted in a more than 50% increase in glycogen content, hence the term muscle glycogen supercompensation. More recent studies have shown that well-trained athletes are able to achieve similar muscle glycogen supercompensation without the need to undertake a glycogen-depletion phase (see Kiens, 2001 for review). Nowadays, the more practiced method for glycogen loading in human athletes involves 3 days of exercise taper combined with a high carbohydrate intake (7-10 g/kg bwt per day). In horses, there also is evidence muscle glycogen content affects exercise performance (Lacombe et al., 2001). However, unlike humans, the phenomenon of glycogen supercompensation does not occur in horses (Lacombe et al., 2004) and only very modest (~10%) increases in muscle glycogen content can be achieved through dietary manipulation (i.e. an increase in dietary starch) (Topliff et al. 1985; Pagan et al. 1987; Essen-Gustavsson et al., 1991). Changes of this magnitude are not associated with improved performance of horses during moderate or high intensity exercise. It is also worth noting that in humans a daily carbohydrate intake of 7-10 g/kg bwt is required for a substantial increase in muscle glycogen content. For a 500-kg horse, an equivalent dose of hydrolyzable carbohydrate would require the consumption of 7 to 10 kg of oats (~50% starch) per day. This level of grain intake is not recommended given possible risk of hindgut disturbance and colic. Fat Adaptation There is widespread belief that adaptation to a fat- or oil-supplemented diet is associated with improvement in the athletic performance of horses. A number of mechanisms have been put forth as potential explanations for improved performance with adaptation to a fat-supplemented diet, including: 1) an improved power-to-weight ratio due to a reduction in DM intake and bowel ballast (Kronfeld, 1996); 2) decreased metabolic heat production associated with feeding and exercise (Kronfeld, 1996); 3) enhanced stamina as a result of muscle glycogen sparing (Kronfeld et al., 1998); 4) improved sprint performance due to increased energy transduction from anaerobic glycolysis (Oldham et al., 1990; Kronfeld et al., 1998); and 5) mitigation of acidemia during high intensity exercise (Kronfeld et al., 1998). Fat supplementation in horses is characterized by a dose-dependent increase in the activity of lipoprotein lipase and, in some studies, an increase in the activity of skeletal muscle citrate synthase and beta-hydroxy acyl-CoA dehydrogenase (Orme et al., 1997; Dunnett et al., 2002). These adaptations suggest that horses adapted to a fat-supplemented diet have increased capacity for the uptake and oxidation of fatty acids in muscle. Indeed, horses fed a diet providing approximately 25% of DE from fat have lower respiratory exchange ratio (Dunnett et al., 2002; Pagan et al., 2002) and decreased glucose utilization (Pagan et al., 2002) during low intensity (~25-35% VO2max) exercise when compared to the control diet. Thus, fat-supplementation enhances lipid oxidation and spares the use of endogenous CHO (plasma glucose, muscle glycogen) during low intensity exercise. Theoretically, such a glycogen-sparing effect could enhance the performance, particularly during endurance exercise in which depletion of muscle glycogen stores is one factor that can limit exercise capacity. However, it is noteworthy that fat adaptation does not result in a clear enhancement of exercise capacity or performance in humans (Helge et al., 1996; Burke et al., 2000; Burke and Hawley 2002). In fact, there is evidence of an increase in the perceived effort of training and an impairment of the response to training when the high-fat, lowcarbohydrate diet is maintained for periods longer than 4 weeks (Helge et al., 1996; Burke and Kiens 2006). What was once viewed as a glycogen-sparing effect after adaptations to a high fat diet may actually be a reflection of a down regulation of carbohydrate metabolism (glycogen impairment). One study of human athletes has reported that fat adaptation is associated with a reduction in the activity of skeletal muscle pyruvate dehydrogenase (Stellingwerff et al., 2006). It was suggested that this decrease in pyruvate dehydrogenase activity could impair muscle glycogenolysis at a time when requirements for carbohydrate are high. The implications of these findings to equine sports nutrition are not known. Several studies have examined the effects of fat supplementation on exercise capacity or performance of horses. Some of these studies have been conducted under field conditions (e.g. a

21

simulated race) while others have employed treadmill exercise protocols (e.g. run time to volitional fatigue during low- or high-intensity treadmill exercise) (Eaton et al., 1995; Essen-Gustavsson et al., 1991; Harkins et al., 1992; Meyers et al., 1989; Oldham et al., 1990; Pagan et al., 1987; Scott et al., 1992; Topliff et al., 1985; Webb et al., 1987). Some authors have reported improved performance (Eaton et al., 1995; Harkins et al., 1992; Hyyppa et al., 1999; Meyers et al., 1989; Oldham et al., 1990; Scott et al., 1992; Webb et al., 1987) while others have found no change (Essen-Gustavsson et al., 1991; Hyyppa et al., 1999; Pagan et al., 1987). A number of factors could account for the variability in results between studies, including the type (e.g. animal vs. vegetable sources) and amount of fat supplemented, the duration of fat feeding, variation in the intensity and duration of exercise tests, the small number of horses per treatment (most often less than 6), and differences in the physical conditioning status of the horses. Harkins et al. (1992) reported that 14 of 15 horses ran a 1600 m simulated race faster when fed a corn oil-supplemented diet (3% DM for 3 weeks) compared to a control ration. Mean race time improved 2.5 sec (2.1%) after consuming the fat-added diet, mainly due to increased speed over the first 200 m. This study by Harkins and colleagues (1992) is often cited as evidence for the ergogenic effects of fat supplementation. However, the design of this experiment was not ideal and its limitations warrant mention. First, the experiment was run in a longitudinal fashion with all horses completing the controldiet race first, with the fat-diet race undertaken after a further 3 weeks of training. It is possible, therefore, that the observed decrease in simulated race time was due to training rather than fat (corn oil) supplementation. Second, the horses were fed hay in the control period but not when the oil-supplemented diet was fed. As a result, the horses received a larger percentage of digestible energy from starch, which could explain the higher muscle glycogen concentrations reported after the period of oil supplementation. Alternatively, it is possible the faster race time was associated with a reduction in gut weight (bowel ballast) due to decreased forage intake. In summary, although there is inconsistency in results there is some evidence that fat supplementation of horses in training may enhance performance during exercise requiring single or repeated, high-intensity efforts. Enhancement in the capacity for fat oxidation following fat adaptation also may be beneficial for endurance-type exercise. However, the adverse effects of high-fat diets on human endurance performance point to the need for well-designed studies that directly assess the effects of fat supplementation on the endurance capacity of horses. For the Future? Temporary or permanent loss of use due to skeletal (bone and/or cartilage) injury is the most important cause of economic loss in the equine industry. In some situations these injuries are catastrophic (e.g. breakdown injuries in racehorses), requiring euthanasia of the affected animal. Increasingly, concerns are being raised about the animal welfare implications of injuries sustained by horses, particularly those engaged in high profile disciplines such as Thoroughbred racing (Bailey, 1998). Is there a role for nutrition in mitigation of skeletal injuries? In humans, low birth weight may have adverse effects on adult bone mineral density and bone mineral content (Cooper et al., 2002), and research is now focused on the relationship between early development (including the intra-uterine environment) and risk of adult disease. Of particular interest are the findings of Verheyen et al. (2006) regarding the effects of dam age and parity on the rate of fracture in offspring in Thoroughbred racehorses in training for flat racing. These authors reported that foals from maiden mares (i.e. first foals) had a significantly lower fracture than foals born to multiparous mares, and the rate of fracture decreased with increasing dam age. These data raise questions concerning the effects of the intra-uterine environment on skeletal development and risk for injury later in life. Our knowledge of how diet during pregnancy affects skeletal development of the foal (both pre- and post-birth) is limited. This area of research may bear the most fruit in the context of advances in feeding management of the equine athlete. Indeed, from a welfare and economic perspective, our goal should be development of management practices, including nutrition, that reduce disease and injury rather than boost the performance of athletic horses.

22

References Andrews, F.M., B.R. Buchanan, S.B. Elliott, N.A. Clariday and L.H. Edwards. 2005. Gastric ulcers in horses. J. Anim. Sci. 83:E18-E21. Bailey, C.J. 1998. Wastage in the Australian Thoroughbred racing industry. RIRDC Equine R & D Program Research Paper No. 98/52. Bergstrom, J., L. Hermansen, E. Hultman and B. Saltin. 1967. Diet, muscle glycogen and physical performance. Acta Physiol. Scand. 71:140-150. Burke, L.M. 2000. Preparation for competition. Pages 341-368. In: Clinical Sports Nutrition. Burke, L., and V. Deakin (eds.) McGraw-Hill, Roseville, NSW, Australia. Burke, L.M., D.J. Angus, G.R. Cox, N.L Cummings, M.A. Febbraio, K. Gawthron, J.A. Haley, M. Minehan, D.T. Martin and M. Hargreaves. 2000. Effects of fat adaptation and carbohydrate restoration on metabolism and performance during prolonged cycling. J.Appl. Physiol. 80:2413-2421. Burke, L.M., and J.A. Hawley. 2002. Effects of short-term fat adaptation on metabolism and performance of prolonged exercise. Med. Sci. Sports Exer. 34:1492-1498. Burke, L.M., and B. Kiens. 2006. Fat adaptation for athletic performance: The nail in the coffin? J. Appl. Physiol. 100:7-8. Cooper, C., M.K. Javaid, P. Taylor, K. Walker-Boone, E. Dennison and N. Arden. 2002. The fetal origins of osteoporotic fracture. Calcif. Tissue Internat. 70:391-394. Crandell, K. 2002. Trends in feeding the American Endurance horse. Pages 135-139. In: Proc. 2002 Kentucky Equine Research, Inc., Equine Nutrition Conference. Cuddeford, D. 2001. Starch digestion in the horse. Pages 95-103. In: Advances in Equine Nutrition II. Pagan, J.D., and R.J. Geor (eds.). Nottingham University Press, Loughborough. Dunnett, C., D.J. Marlin and R.C. Harris. 2002. Effect of dietary lipid on response to exercise: relationship to metabolic adaptation. Eq. Vet. J. Suppl. 34:75-80. Eaton, M.D., D.R. Hodgson, D.L. Evans, W.L. Bryden and R.J. Rose. 1995. Effect of a diet containing supplemental fat on the capacity for high intensity exercise. Eq. Vet. J. Suppl. 18:353-356. Essen-Gustavsson, B., E. Blomstrand, K. Karlstrom and S.G.B. Persson. 1991. Influence of diet on substrate metabolism during exercise. Pages 288-298. In: Equine Exercise Physiology 3. Persson, S., A. Lindholm and L.B. Jeffcott (eds.). ICEEP Publications, Davis, CA. Gallagher, K., J. Leech and H. Stowe. 1988. Protein, energy and dry matter consumption by racing Thoroughbreds: A field study. J. Eq. Vet. Sci. 12:43-47. Geor, R.J. 2006. The role of nutritional supplements and feeding strategies in equine athletic performance. Eq. Comp. Exerc. Physiol. 3:109-119. Harkins, J.D., G.S. Morris, R.T. Tulley, A.G. Nelson and S.G. Kamerling. 1992. Effect of added dietary fat on racing performance in Thoroughbred horses. J. Eq. Vet. Sci. 12:123-129. Harris, P.A. 1999. Review of equine feeding and stable management practices in the UK concentrating on the last decade of the 20th century. Eq. Vet. J. Suppl. 28:46-54. Harris, P.A., and R.C. Harris. 2005. Ergogenic potential of nutritional strategies and substances in the horse. Livestock Prod.. Sci 95:147-165. Hawley, J., E.J. Schabort and T.D. Noakes. 1997. Carbohydrate-loading and exercise performance. An update. Sports Med. 24:73-81. Helge, J.W., E.A. Richter and B. Kiens. 1996. Interaction of training and diet on metabolism and endurance during exercise in man. J. Physiol. 492:293-306. Hinchcliff, K.W., and R.J. Geor. 2004. Integrative physiology of exercise. Pages 3-8. In: Equine Sports Medicine and Surgery. Hinchcliff, K.W., A.J. Kaneps and R.J. Geor (eds.). Elsevier, Edinburgh, UK. Hudson, J.M., N.D. Cohen, P.G. Gibbs and J.A. Thompson. 2001. Feeding practices associated with colic in horses. J. Am. Vet. Med. Assoc. 219:1419-1425. Hyypp, S., M. Saastamoinen and A.R. Ps. 1999. Effect of a post-exercise fat-supplemented diet on muscle glycogen repletion. Eq. Vet. J. Suppl. 30:493-498.

23

Kiens, B. 2001. Diet and training in the week before competition. Can. J. Appl. Physiol. Suppl. 26:S56S63. Kronfeld, D.S. 1996. Dietary fat affects heat production and other variables of equine performance under hot and humid conditions. Eq. Vet. J. Suppl. 22:24-34. Kronfeld, D.S., S.E. Custalow, P.L. Ferrante, L.E. Taylor, J.A. Wilson and W. Tiegs. 1998. Acid-base responses of fat-adapted horses: relevance to hard work in the heat. Appl. Anim. Behav. 59:61-72. Lacombe, V.A., K.W. Hinchcliff, R.J. Geor and C.R. Baskin. 2001. Muscle glycogen depletion and subsequent replenishment affect anaerobic capacity of horses. J. Appl. Physiol. 91:1782-1790. Lacombe, V.A., K.W. Hinchcliff, C.W. Kohn, S.T. Devor and L.E. Taylor. 2004. Effects of feeding meals with various soluble carbohydrate content on muscle glycogen synthesis after exercise in horses. Am. J. Vet. Res. 65:916-923. Mtayer, N., M. Lhte, A. Bahr, N.D. Cohen, I. Kim, A.J. Roussel and V. Julliand. 2004. Meal size and starch content affect gastric emptying in horses. Eq. Vet. J. 36:436-440. Meyer, H., S. Radicke, E. Kienzle, S. Wilke and D. Kleffen. 1993. Investigations on preileal digestion of oats, corn and barley starch in relation to grain processing. Pages 93-97. In: Proc. 13th Equine Nutr. Physiol. Soc. Meyers, M.C., G.D. Potter, J.W. Evans, L.W. Greene and S.F. Crouse 1989. Physiologic and metabolic responses of exercising horses fed added dietary fat. J. Eq. Vet. Sci. 9:218-223. National Research Council. 2007. Nutrient Requirements of Horses, 6th ed. National Academy Press, Washington, DC. Oldham, S.L., G.D. Potter, J.W. Evans, S.B. Smith, T.S. Taylor and W.S. Barnes. 1990. Storage and mobilization of muscle glycogen in exercising horses fed a fat-supplemented diet. J. Eq. Vet. Sci. 10:353-359. Orme, C.E., R.C. Harris, D.J. Marlin and J.S. Hurley. 1997. Metabolic adaptation to a fat supplemented diet in the thoroughbred horse. Brit. J. Nutr. 78:443-458. Pagan, J.D., B. Essen-Gustavsson, A. Lindholm and J. Thornton. 1987. The effect of dietary energy source on exercise performance in Standardbred horses. Pages 686-700. In: Equine Exercise Physiology 2. Robinson, N. (ed.). ICEEP Publishing, Davis, CA. Pagan, J.D., R.J. Geor, P.A. Harris, K. Hoekstra, S. Gardner, C. Hudson and A. Prince. 2002. Effects of fat adaptation on glucose kinetics and substrate oxidation during low intensity exercise. Eq. Vet. J Suppl 34:33-38. Radicke, S., E. Kienzle and H. Meyer. 1991. Preileal apparent digestibility of oats and cornstarch and consequences for cecal metabolism. Pages 43-48. In: Proc. 12th Eq. Nutr. Physiol. Soc. Symp. Scott, B.D., G.D. Potter, L.W. Greene, P.S. Hargis and J.G. Anderson. 1992. Efficacy of a fatsupplemented diet on muscle glycogen concentration in exercising Thoroughbred horses maintained in various body conditions. J. Eq. Vet. Sci. 12:109-113. Southwood, L., D.L. Evans, W.L. Bryden and R.J. Rose. 1993. Nutrient intake of horses in Thoroughbred and Standardbred stables. Aust. Vet. J. 70:164-168. Stellingwerff, T., L.L. Spriet, M.J. Watt, N.E. Kimber, M. Hargreaves, J.A. Hawley L.M. and Burke. Decreased PDH activation and glycogenolysis during exercise following fat adaptation with carbohydrate restoration. Amer. J. Physiol. Endo. Meta. 290:E380-E388. Tinker, M.K., N.A. White, P. Lessard, C.D. Thatcher, K.D. Pelzer, B. Davies and D.K. Carmel. 1997. Retrospective study of equine colic risk factors. Eq. Vet. J. 29:454-458. Topliff, D.R., G.D. Potter, J.L. Kreider, T.R. Dutson and G.T. Jessup. 1985. Diet manipulation, muscle glycogen metabolism, and anaerobic work performance in the equine. Pages 119-124. In: Proc. 9th Conf. Eq. Nutr. Physiol. Soc. Valberg, S.J. 2006. Exertional rhabdomyolysis. Proc. 52nd Conf. Am. Assoc. Eq. Pract. 52:265-372. Verheyen, K.L.P., J.S. Price and J.L.N. Wood. 2006. Fracture rate in Thoroughbred horses is affected by dam age and parity. The Vet. J., doi:10.1016/j.tvjl.2006.07.023. Webb, S.P., G.D. Potter and J.W. Evans. 1987. Physiologic and metabolic response of race and cutting horses to added dietary fat. Pages 115-118. Proc. 10th Conf. Eq. Nutr. Physiol. Soc.

24

BIOFUELS AND BROILERS --- COMPETITORS OR COOPERATORS? Park W. Waldroup Poultry Science Department, University of Arkansas Fayetteville, AR 72701 Phone: 479-575-2065 Email: waldroup@uark.edu Summary The rapid increase in production of ethanol from corn and other grains and biodiesel from various fats and oils has resulted in growing quantities of byproducts from these industries, primarily distillers dried grains with solubles (DDGS) from the ethanol industry and glycerol from biodiesel production. The use of DDGS in poultry and swine diets is not new, but the supply of product encourages the use of higher percentages than has typically been used in the past. As greater quantities are used in the diet, it becomes increasingly essential that accurate nutrient values be assigned to the product. This review attempts to summarize results from various laboratories to provide a nutrient matrix that can be used to evaluate the potential use of DDGS in poultry feeds. Glycerol from biodiesel production has had limited use in animal feeds but results from recent feeding trials indicate that it can be useful as a pure energy source at levels up to 5% of the diet. Introduction DISTILLERS DRIED GRAINS WITH SOLUBLES Byproducts of the distilling industry such as distillers dried grains and distillers dried grains with solubles (DDGS) have long been commonly accepted feed ingredients in broiler diets. Early studies have been extensively reviewed by Scott (1965, 1970). Due to their supply and price, these products were typically fed at levels not exceeding 5% of the diet. However, early studies demonstrated that higher levels could be used in nutritionally balanced diets. Runnels (1966; 1968) reported that 20% DDGS was successfully incorporated into broiler diets with performance equal or superior to that of chicks fed diets with corn, soybean meal, and fish meal. Waldroup et al. (1981) reported that when DDGS was included into broiler diets with the ME content held constant, up to 25% DDGS could be used without reduction in body weight or feed utilization. When included in diets in which the energy content was allowed to decline as the level of DDGS was increased, there was a decline in performance at DDGS levels of 15% or more. Potter (1966) found that isonitrogenous diets with 20% DDGS supported performance equivalent to control diets when fed to poults up to 8 wk of age. Couch et al. (1970) reported that up to 37% DDGS could be used in diets for broiler breeder replacements. Jensen (1978) reported that 20% DDGS was acceptable in nutritionally balanced layer diets and contributed a factor that improved Haugh Units of eggs. Jensen (1981) stated that 20% DDGS could be used in diets for broiler breeder hens. If the modern DDGS from fuel ethanol production is equal or superior in nutritional value to the old DDGS, it is reasonable to expect that satisfactory performance can be obtained when reasonable levels are included in nutritionally-adequate diets. Lumpkins et al. (2004) indicated that DDGS from modern ethanol plants can be safely used at 6% in broiler starter diets and 12 to 15% in grower and finisher periods. Lumpkins et al. (2005) suggested a maximal inclusion rate of 10-12% DDGS in diets for laying hens. Roberson et al. (2005) reported that 15% DDGS did not adversely affect performance of laying hens but suggested that lower levels of DDGS be used when introducing it into the diet. Swiatkiewicz and Korleski (2006) reported that up to 15% DDGS could be used in layer feeds; inclusion of 20% negatively affected laying rate and egg weight. Roberson (2003) noted that DDGS could be effectively included at 10% in growing-finishing diets for turkey hens if proper formulation matrix values for all nutrients were used. Noll and Brannon (2006) reported that performance of turkeys fed 20% DDGS
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

25

was not different from the corn-soybean control unless used in combination with high levels of poultry byproduct meal (8-12%). Like any byproduct, several concerns exist regarding the use of DDGS in poultry feed. These relate primarily to the extent of overall nutrient variability. Major concerns include 1) variation in metabolizable energy content; 2) content and bioavailability of lysine; 3) content and bioavailability of phosphorus, and 4) variation in sodium content. In order to properly utilize DDGS, accurate information regarding the nutrient values for the specific product available is essential. While a general knowledge of the average nutrient content of DDGS in general is helpful, the extreme variability that has been observed in various studies raises a great deal of concern as higher usage levels are contemplated. These issues are addressed below. Variable nutrient content Reliable nutrient values for DDGS are important for optimum use of this product in swine and poultry diets, and recent studies have provided information on various nutrients such as amino acid, metabolizable energy, and mineral contents of DDGS from new ethanol plants (Table 1). Table 1. Proximate composition and amino acid content of DDGS (%, as fed basis). 1 n=118 Mean CV2 88.90 1.7 26.85 6.4 9.69 7.8 7.82 8.7 5.15 14.7 1.07 9.1 0.68 7.8 1.00 8.7 3.16 6.4 0.76 17.3 0.49 13.6 1.31 1.00 0.22 1.33 6.6 6.4 6.7 7.2 Reference1 2 3 4 n=150 n=20 n=5 Mean SD Mean CV Mean 89.91 1.71 88.00 0.9 26.05 2.32 9.88 2.80 14.00 4.8 6.34 1.55 4.39 0.87 4.00 5.0 1.11 0.13 1.00 0.65 0.92 0.18 0.98 2.87 0.63 3.07 0.71 0.17 0.73 11.6 0.64 0.50 0.12 0.49 9.7 0.48 0.54 0.10 0.52 11.3 1.34 0.93 0.17 0.98 6.0 0.95 0.21 0.03 0.25 1.27 0.22 1.30 1.04 5 n=8 Mean SD 28.12

Component Dry matter Crude protein Fat Fiber Ash Arginine Histidine Isoleucine Leucine Lysine Methionine Cystine Phenylalanine Threonine Tryptophan Valine Serine
1

1.09 0.69 0.97 3.05 0.71 0.54 0.56 1.31 0.96 0.20 1.33 1.09

0.16 0.06 0.06 0.14 0.16 0.06 0.04 0.04 0.06 0.05 0.07 0.07

Weighted average 89.36 26.45 10.08 6.99 4.67 1.09 0.68 0.96 3.00 0.73 0.50 0.54 1.31 0.96 0.21 1.30 1.07

1=Spiehs et al. (2002); 2=Fiene et al. (2006); 3=Parsons et al. (2006); 4=Fastinger et al. (2006); 5=Batal and Dale (2006). 2 Coefficient of variation.

It is apparent that there is considerable variability in many of the essential nutrients. Because corn itself varies in nutrient content, concentrating these nutrients approximately three-fold exacerbates the variability in the residual DDGS. For example, Reese and Lewis (1989) reported that corn produced in Nebraska in 1988 ranged from 7.8 to 10.0 % crude protein, 0.22 to 0.32% lysine, and 0.24 to 0.34% phosphorus. In addition, the ratio of blending the distillers solubles with the residual grains to produce DDGS may vary among producers. Some producers add all of the solubles back, while some divert a portion for other uses including use as a fuel source for the ethanol plant. For most of the major nutrients,

26

Spiehs et al. (2002) reported almost as much variation within a source as between different plants. Thus, a continual quality control program to characterize the product will be essential if optimum usage is to be made of DDGS in a poultry formula. Amino acid content and lysine digestibility Extensive analyses of the amino acid contents of DDGS samples have recently been reported by several investigators (Table 1). Mean values for lysine and methionine, the two most critical amino acids for poultry and swine, were similar, but considerable variability was observed. Fiene et al. (2006) conducted stepwise regression analysis on data from approximately 150 samples to predict total amino acid content from the proximate values of moisture, crude protein, fat, and fiber and reported the following equations and R2 values (in parentheses): Arginine: Isoleucine: Leucine: Lysine: Methionine: Cystine: TSAA: Threonine: Tryptophan: Valine: Y = 0.07926 + 0.0398*CP Y= -0.23961 + 0.04084*CP + 0.01227*Fat Y= -1.15573 + 0.13082*CP + 0.06983*Fat Y= -0.41534 + 0.04177*CP + 0.00913*Fiber Y= -0.17997 + 0.02167*CP + 0.01299*Fat Y = 0.11159 + 0.01610*CP + 9.00244*Fat Y= -0.12987 + 0.03499*CP + 0.05344*Fat 0.00229*Fat2 Y= -0.05630 + 0.03343*CP + 0.02989*Fat 0.00141*Fat2 Y= 0.01676 + 0.0073*CP Y= 0.01237 + 0.04731*CP + 0.00054185*Fat2 (0.48) (0.86) (0.86) (0.45) (0.78) (0.52) (0.76) (0.87) (0.31) (0.81)

The R2 values suggest that some amino acids (Ile, Leu, Met, TSAA, Thr, and Val) could be predicted with some success from the proximate values. However, others such as Arg, Cys, Lys, and Trp could not be predicted with a high degree of accuracy due largely to a lack of consistency of the amino acid to protein ratio in the samples tested. Nutritionists are concerned not only with total amino acid content but also the digestibility. Of greatest concern with DDGS is the bioavailability of lysine, as during the process of drying DDGS the material is typically exposed to temperatures of approximately 315 C (600 F). The adverse effect of excess heat on amino acid availability and especially on lysine is well known (McGinnis and Evans, 1947; Warnick and Anderson, 1968). Several recent studies have evaluated the digestibility of amino acids in DDGS and the results are summarized in Table 2.

27

Table 2. Digestible amino acid coefficients (%) of DDGS. Reference1 1 n=8 Mean 84.1 84.1 83.3 88.6 69.6 86.8 73.9 87.5 74.5 82.8 79.3 81.9 2 n=47 Mean SD 85.2 3.46 81.8 89.3 65.9 86.1 77.6 74.6 83.9 81.8 3.56 2.49 9.50 2.70 4.98 4.15 5.08 2.85 3 n=20 Mean CV2 4 n=5 Mean 88.3 85.3 84.1 90.2 76.5 88.5 81.6 88.0 77.5 88.2 81.4 84.3 Weighted average 85.3 84.5 82.2 89.3 68.5 86.8 77.3 87.7 75.1 84.1 81.4 82.8

Amino acid Arginine Histidine Isoleucine Leucine Lysine Methionine Cystine Phenylalanine Threonine Tryptophan Valine Serine
1 2

SD 6.6 5.7 4.9 2.0 11.5 3.4 9.7 3.3 6.0 5.1 3.3 4.3

72 88 77 76

11.2 1.9 7.7 4.8

1=Batal and Dale (2006); 2=Fiene et al. (2006); 3=Parsons et al. (2006); 4=Fastinger et al. (2006). Coefficient of variation.

The digestibility of lysine is the lowest among the amino acids and also has the greatest variability. A rapid means of assessing the lysine digestibility in a particular sample of DDGS is of prime importance in optimizing the use of this ingredient, and two methods recently have been proposed in this regard. One of these is the use of the Immobilized Digestibility Enzyme Assay (IDEA, Novus International, St. Louis MO) described by Shasteen et al. (2002). This assay was used to estimate the lysine digestibility of 28 DDGS samples that had previously been subjected to in vivo-determined true lysine digestibility. There was a high correlation between in vivo-determined true lysine digestibility and that estimated by the IDEA method (Fiene et al. 2006). The correlation between digestibility of other amino acids and the IDEA method was not as successful, ranging from 0.12 for Met to 0.43 for Cys. More than 180 samples of DDGS have been subjected to the IDEA assay by Novus International, resulting in an estimated lysine digestibility of 66.7 9.3 (mean SD). This is in good agreement with the weighted average of 68.5% shown in Table 2. In order to evaluate variation among producers, multiple DDGS samples were collected from eight different suppliers over a 3 to 4 month time period and subjected to the IDEA assay (Fiene et al., 2006). The data showed that within a supplier the variation in lysine digestibility was relatively small with a few exceptions. It is recommended that the IDEA assay be used periodically to estimate the digestible lysine content of samples received in a feed mill, especially since the product may come from a wide variety of sources. A second method that has been used to estimate lysine digestibility is evaluation of the color of the product. Formation of lysine-carbohydrate complexes under heat has long been known (Maillard, 1912a,b). Color of soybean meal has long been linked to proper processing temperature (McNaughton et al., 1981). Numerous studies have linked the color of DDGS with lysine digestibility (Cromwell et al., 1993; Ergul et al., 2003; Batal and Dale, 2006; Fastinger et al., 2006). Batal and Dale (2006) reported that samples with more lightness (L*=60.3) and more yellowness (b*=25.9) were associated with a DDGS having an average of 0.66% digestible Lys whereas products that were darker (L*=50.4) and less yellow (b*=7.41) were associated with a product having 0.18% digestible Lys. Use of visual color or use of color meters may be used to identify samples of DDGS that have been subjected to excessive heat with subsequent reduction in lysine bioavailability.

28

Metabolizable energy content of DDGS Several studies provide estimates of the metabolizable energy content of DDGS. Batal and Dale (2006) reported an average TMEn for 17 samples of 1282 82 kcal/lb, with a range of 1132 to 1450 kcal/lb. Fastinger et al. (2006) found an average of 1302 kcal/lb for five samples with a range of 1127 to 1382 kcal/lb. Lumpkins and Batal (2005) reported a TMEn value of 1318 kcal/lb for a single sample of DDGS. Parsons et al. (2006) reported an average TMEn of 1299 kcal/lb for 20 samples with a range from 1182 to 1385 kcal/lb. A weighted average of these 43 samples is 1293 kcal/lb. Batal and Dale (2006) applied regression analyses to the proximate composition of the DDGS and the determined TMEn values and developed the following equations (R2 values in parentheses), which can be applied to samples of DDGS to estimate the TMEn value: TMEn = 2439.4 + 43.2*Fat TMEn = 2957.1 + 43.8*Fat - 79.1*Fiber TMEn = 2582.3 + 36.7*Fat 72.4*Fiber + 14.6*Protein TMEn = 2732.7 + 36.4*Fat -76.3*Fiber + 14.5*Protein 26.2*Ash DDGS mineral content and relative phosphorus bioavailability Several recent reports on the mineral content of DDGS have been published and are summarized below (Table 3). The bioavailability of the P in DDGS appears to be higher than previously assumed (National Research Council, 1994). This may be in part because of the P provided by the residual yeast, which is considered to be highly available, and because the fermentation may release some P from the phytate bond. As early as 1972, Singsen et al. reported that the biological availability of the phosphorus in three composite samples of DDGS was fully equivalent to that in commercial dicalcium phosphate and should be considered as 100% available when formulating poultry diets. Martinez -Amezcua et al. (2004) noted a substantial variability in P bioavailability among nine samples, ranging from 69 to 102% relative to KH2PO4 , and reported that increased heat processing of DDGS may increase the bioavailability of P in DDGS. Lumpkins and Batal (2005) reported that the relative bioavailability of phosphorus in a DDGS sample containing 0.74% total P was 68 and 54% in two different trials. Martinez-Amezcua et al. (2006) found a relative P bioavailability of 62% in a sample of DDGS containing 0.67% total P. The bioavailability was increased by supplementation of the diet with 3% citric acid or with phytase. Sodium is one of the most inexpensive minerals but deficiency states have perhaps more rapid impact on performance of any essential nutrient. With the demand to reduce litter moisture in poultry houses, nutritionists often are pressured to minimize dietary sodium levels. Considerable variation in sodium content of DDGS has been observed (Table 3). Batal and Dale (2003) noted that the source of the extraordinary variability in sodium content of DDGS is not immediately clear, and suggested that nutritionists need to properly characterize the mineral content of the DDGS from respective sources prior to incorporation into balanced diets. However, data from Spiehs et al. (2002) showed considerable inplant variation indicating that it would be difficult to characterize the sodium content of a single plant by a few analyses. It is recommended that frequent sodium assays be made of the product received in feed mills, especially if the sodium from the DDGS is to be considered in meeting the requirements for this nutrient. (0.29) (0.43) (0.44) (0.45)

29

Table 3. Mineral composition of DDGS from various authors (%, as fed basis). Reference1 1 n=118 Mean CV2 0.05 57.2 0.79 11.7 0.84 14.0 0.21 70.5 2 n=12 Mean SD 0.29 0.27 0.68 0.07 0.91 0.11 0.25 0.15 3 n=20 Mean SD 0.73 0.04 4 n-20 Mean CV 0.03 38.4 0.73 5.3 0.11 32.8 Weighted average 0.07 0.77 0.85 0.20

Mineral Calcium Phosphorus Potassium Sodium


1

1=Spiehs et al. (2002); 2=Batal and Dale (2003); 3=Martinez-Amezcua et al. (2004); 4=Parsons et al. (2006). 2 Coefficient of variation.

Nutrient matrix for DDGS A number of studies have reported on the nutrient content of various samples of DDGS, many of which have been cited in this report. We have summarized these as weighted averages for various nutrients and combined these into a suggested nutrient matrix to use as a starting point for evaluating DDGS in poultry feed (Table 4.). Table 4. Suggested nutrient matrix for DDGS based on weighted averages of published data. Nutrient Dry matter Crude protein Fat Fiber TMEn (kcal/lb) Arginine Histidine Isoleucine Leucine Lysine Methionine Cystine Phenylalanine Threonine Tryptophan Valine Serine Amount (%) 89.36 26.45 10.08 6.99 1293 1.09 0.68 0.96 3.00 0.73 0.50 0.54 1.31 0.96 0.21 1.30 1.07 Nutrient Calcium Phosphorus Available phosphorus Potassium Sodium Digestible Arginine Digestible Histidine Digestible Isoleucine Digestible Leucine Digestible Lysine Digestible Methionine Digestible Cystine Digestible Phenylalanine Digestible Threonine Digestible Tryptophan Digestible Valine Digestible Serine Amount (%) 0.07 0.77 0.48 0.85 0.20 0.93 0.58 0.78 2.70 0.50 0.43 0.42 1.15 0.72 0.18 1.05 0.88

Nutritionists should continuously scrutinize the proximate composition of the product along with periodic assays for calcium, phosphorus and sodium and for estimates of lysine digestibility by IDEA, color meter, or visual inspection of color.

30

Future products may differ in composition As ethanol production from corn increases, there is growing interest in modifying the technology used to produce the product. This will result in different types of byproducts that may have superior or inferior nutritional value (Parsons et al., 2006). Use of these new manufacturing processes will result in the production of byproducts that will undoubtedly differ markedly in nutrient content from those produced today. It will be necessary for the nutritionist to be sure they have accurate nutritional values for the products that they will be using in their diets. Ethanol producers should work with the feed industry to provide characteristic nutrient values for such new products as they develop. GLYCEROL IN BROILER FEEDS An increasing amount of inedible fats and oils are being processed for use as biodiesel. A byproduct of biodiesel production is glycerol, the carbohydrate fraction that makes up about 10-11% by weight of typical triglycerides. Several studies have evaluated this product in diets for poultry and swine (Bernal et al., 1978; Barteczko and Kaminski, 1999; Kijora et al., 1995, 1997; Simon, 1996; Kuhn, 1996; Kijora, 1996; Francois, 1994; Wagner, 1994; Simon et al., 1996, 1997). In the majority of these trials the glycerol used was a byproduct of the conversion of rapeseed oil to biodiesel. Studies have recently been conducted in our laboratory to evaluate glycerol produced from biodiesel plants in the United States. A supply of glycerol was obtained from a biodiesel producer (Griffin Industries, Cold Spring, KY). This product was stated to contain < 0.5% methanol. In a preliminary study in which diets were fed from 1 to 16 d of age, we found that up to 10% glycerol could be included without adversely affecting performance. A second study was then conducted in which broilers were grown to 42 d of age on diets with 0, 5, or 10% glycerol. A metabolizable energy value of 1600 kcal/lb was used for the glycerol, based on bomb calorimetry values of 1636 kcal/lb. Each diet was fed to 8 pens of 60 males. Results of the study (Table 5) demonstrated that diets with 5% glycerol supported performance equivalent to that of positive control diets. Diets with 10% glycerol did not flow well in the tube feeders and inhibited feed intake, resulting in slower growth and poor feed conversion. Litter from pens fed 10% glycerol was visibly wetter and upon analysis the diets contained about 0.15% higher K levels as a result of residual K in the glycerol. In a third study, using glycerol from a second producer (Patriot Biofuels, Stuttgart, AR) levels of 0, 2.5 and 5% glycerol were compared with each diet fed to 8 pens of 60 males from 1 to 42 d of age. Results of this study (Table 5) indicated that performance of broilers fed diets with 2.5 or 5% glycerol did not differ significantly from that of birds fed the control diet. Breast meat yield was significantly improved by the addition of glycerol. Glycerol appears to have promise as a feed ingredient for broilers as a pure energy source. Levels up to 10% appear to be utilized by broilers but may cause problems with feed flow and result in reduced performance. More needs to be learned about quality factors related to glycerol from biodiesel production before extensive use can be recommended in broiler feeds. Specific concerns relate to residual levels of methanol, sodium or potassium, and moisture content. Means of handling glycerol in a feed mill must also be considered.

31

Table 5. Effect of glycerol levels on broiler performance. Trial 1 Measurement 42 d BW (kg) 0-42 d FCR (feed:gain) 0-42 d mortality (%) Dressing percentage (%) Breast (% of carcass weight) 0 2.871a 1.732a 6.45 72.85a 26.45 Trial 2 Glycerol (%) 5 2.879a 1.709a 4.58 72.81a 26.72 10 2.706a 1.786b 5.41 72.17b 25.96

Glycerol (%) 0 2.5 42 d BW (kg) 2.618 2.712 0-42 d FCR (feed:gain) 1.643 1.625 0-42 d mortality (%) 6.25 7.50 Dressing percentage (%) 72.05 72.34 Breast (% of carcass weight) 25.16b 25.80a a,b Within rows, means with similar superscripts do not differ significantly (P>0.05) References

5 2.709 1.629 6.45 72.08 25.96a

Barteczko, J., and J. Kaminski. 1999. The effect of glycerol and vegetable fat on some physiological indices of the blood and excess of fat in broiler carcasses. Ann. Warsaw Agric. Univ. Anim. Sci. 36:197-209. Batal, A., and N. Dale. 2003. Mineral composition of distillers dried grains with solubles. J. Appl. Poult. Res. 12:400-403. Batal, A.B., and N.M. Dale. 2006. True metabolizable energy and amino acid digestibility of distillers dried grains with solubles. J. Appl. Poult. Res. 15:89-93. Bernal, G., J.D. Garza, M. Viana, E. Avila, A.S. Shimada, and M. Montano. 1978. Effect of inclusion of glycerol or vegetable oil in diets with molasses for growing pigs and poultry. Vet. Mex. 9:91-94. Couch, J.R., J.H. Trammell, A. Tolan, and W.W. Abbott. 1970. Corn distillers dried grains with solubles in low lysine diets for rearing broiler breeder replacement pullets. Proc. Distillers Res. Council 25:2533. Cincinnati, OH. Cromwell, G.L., K.L. Herkelman, and T.S. Stahly. 1993. Physical, chemical, and nutritional characteristics of distillers dried grains with solubles for chicks and pigs. J. Anim. Sci. 71:679-686. Ergul, T., C. Martinez-Amezcua, C.M. Parson, B. Walters, J. Brannon, and S.L. Noll, 2003. Amino acid digestibility in corn distillers dried grains with solubles. Poultry Sci. 82(Suppl.1):70. Fastinger, N.D., J.D. Latshaw, and D.C. Mahan. 2006. Amino acid availability and true metabolizable energy content of corn distillers dried grains with solubles in adult cecectomized roosters. Poult. Sci. 85:1212-1216. Feine, S.P., T.W. York, and C. Shasteen, 2006. Correlation of DDGS IDEA digestibility assay for poultry with cockerel true amino acid digestibility. Pages 82-89. In: Proc. 4th Mid-Atlantic Nutrition Conference. University of Maryland, College Park, MD. Francois, A. 1994. Glycerol in nutrition. C.R. Acad. Agric. Fr. 80(2):63-76. Jensen, L.S. 1978. Distillers feeds as sources of unidentified factors for laying hens. Proc. Distillers Feed Conf. 33:17-22. Louisville, KY Jensen, L.S. 1981. Value of distillers dried grains with solubles in poultry feeds. Proc. Distillers Feed Conf. 36:87-93 Cincinnati, OH. Kijora, C. 1996. Utilization of glycerol as a byproduct of Bio-Diesel production in animal nutrition. Landbauforshung Volkenrode 169:151-157.

32

Kijora, C., H. Bergner, R.D. Kupsch, and L. Hagemann. 1995. Glycerol as a feed component in diets of fattening pigs. Arch. Anim . Nutr. 47:345-360. Kijora, C., R. D. Kupscy, H. Bergner, C. Wenk, and A. L. Prabucki. 1997. Comparative investigation on the utilization of glycerol, free fatty acids, free fatty acids in combination with glycerol and vegetable oil in fattening of pigs. J. Anim. Physiol. Anim. Nutr. 77:127-138. Kuhn, M. 1996. Use of technical rapeseed-glycerol from Bio-Diesel production in the fatting of pigs. Lanbauforshung Volkenrode 169:163-167. Lumpkins, B.S., and A.B. Batal. 2005. The bioavailability of lysine and phosphorus in distillers dried grains with solubles. Poult. Sci. 84:581-586. Lumpkins, B.S., A. B. Batal, and N.M. Dale. 2004. Evaluation of distillers dried grains with solubles as a feed ingredient for broilers. Poult. Sci. 83:1891-1896. Lumpkins, B., A. Batal, and N. Dale, 2005. Use of distillers dried grains plus solubles in laying hen diets. J. Appl. Poult. Res. 14:25-31. Maillard, L.C. 1912a. Action des acides amines sure les sucres: formation des melanoidines par voie methodique. Comp. Rend. Acad. Sci. 154:66. Maillard, L.C. 1912 b. Formation dhumus et de combustibles mineraux sans intervention de loxygene atmosperique des microorganisms, des haute temperatures, ou des fortes precions. Comp. Rend. Acad. Sci. 155:1554. Martinez-Amezcua, C., C.M. Parsons, and D.H. Baker. 2006. Effect of microbial phytase and citric acid on phosphorus bioavailability, apparent metabolizable energy, and amino acid digestibility in distillers dried grains with solubles in chicks. Poult. Sci. 85:470-475. Martinez-Amezcua, C., C.M. Parsons, and S.L. Noll. 2004. Content and relative bioavailability of phosphorus in distillers dried grains with solubles in chicks. Poult. Sci. 83:971-976. McGinnis, J., and R.J. Evans. 1947. Amino acid deficiencies in raw and overheated soybean meal. J. Nutr. 34:725-732. McNaughton, J.L., F.N. Reese, and J.W. Deaton. 1981. Relationship between color, trypsin inhibitor, contents, and urease index of soybean meal and effects on broiler performance. Poult. Sci. 60:393-400. National Research Council, 1994. Nutrient Requirements of Poultry. 9th rev. ed. National Academy Press, Washington, DC. Noll, S.L., and J. Brannon, 2006. Inclusion levels of corn distillers grains with solubles and poultry byproduct meal in market turkey diets. Poult. Sci. 85 (Suppl. 1):106-107. Parsons, C.M., C. Martinez, V. Singh, S. Radhadkrishman, and S. Noll. 2006. Nutritional value of conventional and modified DDGS for poultry. Proc. Multi-State Poult. Nutr. Feeding Conf., Indianapolis, IN. Potter, L.M. 1966. Studies with distillers feeds in turkey rations. Proc. Distillers Res. Conf. 21:47-51. Cincinnati, OH. Reese, D.E., and A.J. Lewis, 1989. Nutrient content of Nebraska corn. Neb. Coop. Ext. Service EC 89219, pp. 5-7. Roberson, K.D. 2003. Use of dried distillers grains with solubles in growing-finishing diets of turkey hens. Int. J. Poult. Sci. 2:389-393. Roberson, K.D., J. L. Kalbfleisch, W. Pan, and R.A. Charbeneau. 2005. Effect of corn distillers dried grains with solubles at various levels on performance of laying hens and egg yolk color. Int. J. Poult. Sci. 4:44-51. Runnels, T.D. 1966. The biological nutrient availability of corn distillers dried grains with solubles in broiler feeds. Proc. Distillers Research Council 21:11-15, Cincinnati, OH. Runnels, T.D. 1968. Effective levels of distillers feeds in poultry rations. Proc. Distillers Res. Council. 23:15-22, Cincinnati, OH. Scott, M.L. 1965. Distillers dried solubles for maximum broiler growth and maximum early egg size. Proc. Distillers Research Council 25:55-57, Cincinnati, OH. Scott, M.L. 1970. Twenty-five years of research on distillers feeds for broilers. Proc. Distillers Research Council 25:19-24, Cincinnati, OH.

33

Shasteen, C., J. Wu, M. Schulz, and C. Parsons. 2002. An enzyme based protein digestibility assay for animal production. In: Proc. Multi-State Poult. Nutr. Feeding Conf., Indianapolis, IN. Simon, A., H. Bergner, and M. Schwabe. 1996. Glycerolfeed ingredient for broiler chickens. Arch. Anim. Nutr. 49:103-112. Simon, A., M. Schwabe, and H. Bergner. 1997. Glycerol supplementation to broiler rations with low crude protein content. Arch. Anim. Nutr. 50:271-282. Simon, A. 1996. Administration of glycerol to broilers in the drinking water. Landbauforshung Volkenrode 169:168-170. Singsen, E.P., L.D. Matterson, and J.J. Tlustohowicz. 1972. The biological availability of phosphorus in distillers dried grains with solubles for poultry. Proc. Distillers Research Council 27:46-49, Cincinnati, OH. Spiehs, M.J., M.H. Whitney, and G.C. Shurson. 2002. Nutrient database for distillers dried grains with solubles produced from new ethanol plants in Minnesota and South Dakota. J. Anim. Sci. 80:26392645. Swiatkiewicz, S., and J. Korleski. 2006. Effect of maize distillers dried grains with solubles and dietary enzyme supplementation on the performance of laying hens. J. Anim. Feed Sci. 15:253-260. Wagner, F. 1994. Glycerol in animal feedinga byproduct of alternative fuel production. Muhle+Mischfuttertechnik 131(44)621-622. Waldroup, P.W., J.A. Owen, B.E. Ramsy, and D.L. Whelchel. 1981. The use of high levels of distillers dried grains plus solubles in broiler diets. Poult. Sci. 60:1479-1484. Warnick, R.E., and J.O. Anderson. 1968. Limiting essential amino acids in soybean meal for growing chickens and effects of heat upon availability of essential amino acids. Poult. Sci. 47:281-28.

34

BIOTECHNOLOGY IN THE BARNYARD - WHAT WILL IT LOOK LIKE IN 2050? Terry D. Etherton Department of Dairy and Animal Science 324 W.L. Henning Building The Pennsylvania State University University Park, PA 16802 Phone: 814-863-3665 FAX: 814-863-6042 Email: tde@psu.edu

Summary
Since the onset of the modern era of biotechnology in 1973, scientists have made impressive strides in developing new agricultural biotechnologies (reviewed in National Research Council [NRC], 1994; Etherton et al., 2003). Biotechnologies that enhance productivity and productive efficiency (feed consumed/unit of output) have been developed and approved for commercial use. Technologies that improve productive efficiency will benefit both producers and consumers because feed provision constitutes a major component (about 70%) of farm expenditures. Advances in biotechnology research have allowed impressive improvements to be made in diagnostic approaches, increasing microbial safety of food and improving animal health (reviewed in Etherton et al., 2003). The application of genomics, or the study of how genes (DNA) are organized and expressed, and bioinformatics in animal agriculture will provide new genetic markers for improved selection of all livestock species. The advent of techniques to propagate animals by nuclear transfer (cloning) offers many important applications to animal agriculture, including reproducing highly desired elite sires and dams. Animals selected for cloning will be of great value because of their increased genetic merit for increased food production, disease resistance, reproductive efficiency, or will be valued because they have been genetically modified to produce organs for transplantation or products with biomedical application (The National Academies, 2004). Biotechnology also offers considerable potential to animal agriculture as a means to reduce nutrients and odors from manure and volume of manure produced. Development and adoption of these biotechnologies will contribute to a more sustainable environment. The discovery and development of new animal biotechnologies are part of a continuum leading to the commercialization of agricultural biotechnology products. In order to enter the marketplace, new agricultural biotechnologies are evaluated rigorously by the appropriate federal regulatory agencies to ensure efficacy, consumer safety, and animal health and well being (FDA, 2006). To benefit agriculture and society, products of biotechnology must be accepted by consumers. Central to consumer acceptance is the need to provide effective population-based education programs to enhance public understanding of the safety and benefits associated with technological advances enabled by agricultural biotechnology. Despite some of the most remarkable advances in biological research, a public discussion still continues about the need for, and safety of, agricultural biotechnology that is fueled by misinformation campaigns funded by animal activists and some consumer activist groups. As we progress towards 2050, the scientific and agricultural communities must be more proactive in developing and delivering biotechnology and agriculture education campaigns for public and policy makers that clearly articulate the merits of current production practices used in animal agriculture. Moreover, the benefits of investing in discovery research that improves animal agriculture must be championed, and the return on this investment clearly communicated. The agricultural community is going to navigate a period over the next few decades during which we will likely witness growing challenges, especially increased regulatory oversight in addition to the misinformation campaigns funded by activist anti-ag biotechnology groups. For the full benefits of agricultural biotechnology to be realized, regulatory policy that evolves must be guided by the scientific evidence base, not vocal anti-ag biotechnology activist groups.
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

35

Human Health: Uses of Biotechnology in Animal Agriculture to Decrease Morbidity and Mortality from Disease Advances in recombinant DNA technology, animal embryology, immunology, and other disciplines give rise to the prospect that animals will become important sources of highly sophisticated biopharmaceuticals and biological products. A number of biopharmaceutical products are being developed in which the ultimate production system will be a genetically modified (GM) animal. In such instances, the gene for the desired molecule is designed and constructed to be expressed only in a specific tissue. Several companies focusing on these activities are targeting the production of specific animal proteins in milk (cattle, sheep, goats, pigs) and in eggs (poultry). These new approaches will allow for more economical production of certain pharmaceuticals that currently are expensive to produce. For example, Roslin Institute researchers (Lillico et al., 2007) recently reported the generation of transgenic hens that synthesized two functional recombinant therapeutic proteins in a tightly regulated tissue-specific manner, without any evidence of transgene silencing after germ-line transmission. Biotechnological production methods also may allow for a more widespread distribution and application of pharmaceuticals, including in developing countries. Another type of biotechnology under development involves using animals to produce donor organs for human transplant. There is a great shortage of transplant organs from human donors. Heart valves from pigs have been used for many years to replace heart valves in humans, and several private-sector companies and numerous university scientists are investigating ways in which to use biotechnology to develop pigs as a source of transplant organs. One key advantage of this approach is that it is possible that genetic engineering can be used to produce rejection-free organs. In each instance just described, a vital basic health need is being fulfilled by the use of animals, and no foreseeable alternatives exist. Food Production: Uses of Biotechnology in Animal Production Feeding Livestock Biotechnology has led to the development of plant varieties with improved qualities including enhanced tolerance of herbicides, and protection against viruses and insect pests, and modifications in nutrient profile. These varieties have been adopted rapidly by American farmers, and the United States accounted for approximately 55% of the global area of transgenic crops in 2005 (up from 30% in 2001) (James, 2005). Genetically modified crops used for livestock feed include corn, soybean, canola, and cotton (cottonseed). Health and safety are priorities in the development of new food and feed products, including those developed through biotechnological means. Evaluation by governmental regulatory agencies is required for each new biotech plant used for feed or food. Scientific studies evaluating feed components derived from GM plants have focused on beef cattle, swine, sheep, fish, lactating dairy cows, and broiler and layer chickens, and have included nutrient composition assessments, digestibility determinations, and animal performance measurements (Alewynse 2000; Beever and Kemp 2000; Faust 2002). Evaluations have shown uniformly that feed components derived from GM plants commercialized thus far are substantively equivalent in terms of nutrient composition and are similar in terms of nutrient digestibility and feeding value. Overall, feed components of GM plants result in growth rates and milk yields not different from those derived from non-genetically enhanced feed sources (Clark and Ipharraguerre 2001; Faust 2002; Flachowsky and Aulrich 2001). Studies have reported that when corn has been altered genetically for protection against the corn borer, under certain growing conditions GM plants can have lower mycotoxin contamination, resulting in safer feed for livestock (Munkvold et al., 1999). In addition, no residues of recombinant DNA or novel proteins have been found in any tissue or organ samples obtained from animals fed GM plants (Deaville and Maddison, 2005; Flachowsky et al., 2005).

36

Metabolic Modifiers Advances in understanding the regulation of nutrient use in agricultural animals have led to the development of technologies referred to as metabolic modifiers. Metabolic modifiers are a group of compounds that modify animal metabolism in specific and directed ways. Metabolic modifiers have the overall effect of improving production, productive efficiency (weight gain or milk yield/unit of feed consumed), improving carcass composition (lean:fat ratio) in growing animals, increasing milk yield in lactating animals, and decreasing animal waste/production unit (NRC, 1994). Two classes of compounds have received major focussomatotropins (STs) and -adrenergic agonists. The most commonly discussed ST is bovine somatotropin (bST), which has been commercially used since 1994 for administration to dairy cows to achieve increased milk yield, improve milk/feed, and decrease animal waste (Etherton and Bauman, 1998; Bauman, 1999). Supplements of -adrenergic agonists to growing animals improve feed utilization and increase rate of weight gain, carcass leanness, and dressing percentage (NRC, 1994). Research has established the mode of action involves changes in endocrine and cellular mechanisms (NRC, 1994). The net effect is that these repartitioning agents improve productive efficiency by modifying specific metabolic signals in a coordinated manner to increase nutrient use for lean tissue. Ractopamine (Paylean) is the only adrenergic agonist currently approved in the United Statesin this instance, for finishing pigs; commercial use began in 2000. Cloning Cloning, a term originally used primarily in horticulture to describe asexually produced progeny, means to make a copy of an individual or, in cellular and molecular biology, groups of identical cells, and replicas of DNA and other molecules. For example, monozygotic twins are clones. Animal cloning in the late 1980s resulted from the transfer of nuclei from blastomeres of early cleavage-stage embryos into enucleated oocytes, and cloning of livestock and laboratory animals has resulted from transferring a nucleus from a somatic cell into an oocyte from which the nucleus has been removed (Wilmut et al., 1997; Westhusin et al., 2001). Somatic cell nuclear transfer also can be used to produce embryonic stem cells, which are undifferentiated, and matched to the recipient for research and therapy that is independent of reproductive cloning of animals. The progeny from cloning using nuclei from either blastomeres or somatic cells are not exact replicas of an individual animal due to cytoplasmic inheritance of mitochondrial DNA from the donor egg, other cytoplasmic factors which may influence "reprogramming" of the genome of the transferred nucleus, and subsequent development of the cloned organism (Jaenisch and Wilmut, 2001). Cloning by nuclear transfer from embryonic blastomeres (Willadsen, 1989) or from a differentiated cell of an adult (Wilmut et al., 1997; Polejaeva et al., 2000; Kuhholzer and Prather, 2000) requires that the introduced nucleus be reprogrammed by the cytoplasm of the egg and direct development of a new embryo, which is then transferred to a recipient mother for development to term. The offspring will be identical to their siblings and to the original donor animal in terms of their nuclear DNA, but will differ in their mitochondrial genes; variances in the manner nuclear genes are expressed are also possible. Although clone is descriptive for multiple approaches for cloning animals, in this article clone is used as a descriptor for somatic cell nuclear transfer. On December 28, 2006, the Food and Drug Administration (FDA) released a draft risk assessment (RA) on whether cloning affects food safety or animal health, and whether food products from livestock should be sold for human consumption. The draft, A Risk-Based Approach to Evaluate Animal Clones and Their Progeny DRAFT (visit http://www.fda.gov/cvm/CloneRiskAssessment.htm) concludes that .the available data has not identified any food consumption risks or subtle hazards in healthy clones of cattle, swine, or goats. Thus, edible products from healthy clones that meet existing

37

requirements for meat and milk in commerce pose no increased food consumption risk(s) relative to comparable products from sexually-derived animals. Publication of the FDA Risk Assessment is an important next step in the process leading to the release the final regulatory guidelines that will allow food from cloned animals to enter the food system. Conservation of the Environment Meeting environmental challenges is one of the major issues facing animal agriculture. Most swine and poultry manure is produced in confinement units for which the nearby land base often is insufficient to accommodate waste in an environmentally sound manner. Animal manure, especially that from swine and poultry, is high in nitrogen (N) (4.7 to 5.1%) and phosphorus (P) (1.6 to 3.0%), both of which can contribute to surface and groundwater pollution. In addition, ammonia and other nitrogenous and sulfurous gasses contribute to poor air quality and offensive odors. Several GM crops have been developed or are being developed to address the environmental issues related to N, P, and total manure excretion and odors (Etherton et al., 2003). Phosphorus content in swine and poultry manure is high because these species consume diets consisting of cereal grains and oilseed meals in which most (60 to 80%) P is bound organically as phytic acid or phytate. Because of the lack of phytase in their digestive tract, nonruminants are unable to degrade phytate, and most P from these feed ingredients is excreted in the feces. In addition, relatively large amounts of inorganic P must be fed to pigs and poultry to meet their P requirements; consequently, fecal P excretion is increased further. Ruminants use phytate quite efficiently because of the abundance of phytase produced by rumen microorganisms. It is exciting that opportunities are now available to decrease P content of manure (reviewed in Knowlton et al., 2004). These new strategies are based on a more accurate interpretation of P requirements (to not over-feed P), more precise diet formulation, and utilization of exogenous phytase or low-phytic acid grains in monogastric diets. The availability of microbial expression systems has made large amounts of the recombinant enzyme available for use in animal feed at relatively low costs (reviewed by Lei and Porres, 2003). Collectively, these strategies can lower P content of manure by 4060% in pigs and poultry and 25 to 40% in ruminants (Knowlton et al., 2004). A Look to the Future The impressive growth in the science of biotechnology and the many resulting products, both biomedical and agricultural, that have been developed for society is one of the most impressive achievements in the history of science. Predicting what scientific discoveries will occur between the present and 2050 will, as always, be more than a bit challenging. Scientific advances will give us a better understanding of how genes work, and how they can be manipulated to achieve an optimal production outcome that benefits both the producer and consumer. Valuable animals that arise from conventional breeding or genetic manipulation can be propagated forever by cloning. I anticipate that we will be able to do large-scale modification of a large number of genes that will further enhance a variety of target production traits, production efficiency, and profitability. The extent to which we can enhance these outcomes will be beyond current predictions. Before we in the agricultural community get carried away anticipating scientific advances in biotechnology over the next 40 years, there are several key points that must be considered and addressed. First, funding for discovery and applied research in agriculture must be increased. Second, the discoveries made require a viable private sector to commercialize new products of biotechnology. This is becoming more challenging for a variety of reasons. The process of moving a product through the regulatory approval process is becoming more complex, costly and lengthy. This growing burden makes it challenging for private sector to recover their investment costs from product sales. This is particularly important for agricultural biotechnologies where the margins on products are much lower than biomedical

38

biotechnology products (using similar scientific methods). Over the past 20 years, a number of companies have withdrawn from developing animal health products. What has not been widely addressed is the cost to society if biotechnological innovation in agriculture is hindered or even stopped. The last point pertains to the activist groups that are actively advocating that the use of biotechnology-derived products be halted. This is an ever-present reality. Many of these groups are well funded (visit Guidestar at http://www.guidestar.org/ to see the IRS returns that all non-profits are required to place in the public domain), and attack animal agriculture on many fronts that range from animal welfare to biotechnology to environmental issues. It is simple to scare the public in 30 seconds; however, we can not educate them about science, agriculture, and biotechnology in 30 seconds. My encouragement to animal agriculture is to passionately engage in developing and implementing consumer education programs that effectively frame the importance of animal agriculture and promote the need for and benefits of biotechnology in the barnyard. References Alewynse, M.G. 2000. Regulation of genetically modified plants in animal feed. FDA Vet. 15:12. Bauman, D.E. 1999. Bovine somatotropin and lactation: From basic science to commercial application. Domest. Anim. Endocrinol. 17:101-116. Beever, D.E., and C.F. Kemp. 2000. Safety issues associated with the DNA in animal feed derived from genetically modified crops. A review of scientific and regulatory procedures. Nutr. Abstr. Rev. B: Livestock Feeds Feeding. 70:175-182. Clark, J.H., and I.R. Ipharraguerre. 2001. Livestock performance: Feeding biotech crops. J. Dairy. Sci. 84:E9-E18. Cummins, J.M. 2001. Cytoplasmic inheritance and its implications for animal biotechnology. Theriogenology 55:1381-1399. Deaville, E.R., and B.C. Maddison. 2005. Detection of transgenic and endogenous plant DNA fragments in the blood, tissues, and digesta of broilers. J. Agric. Food Chem. 53:10268-10275. Etherton, T.D., and D.E. Bauman. 1998. The biology of somatotropin in growth and lactation of domestic animals. Physiol. Rev. 78:745-761. Etherton, T.D., D.E. Bauman, C.W. Beattie, R.D. Bremel, G.L. Cromwell, V. Kapur, G. Varner, M.B. Wheeler and M. Wiedmann. 2003. Biotechnology in Animal Agriculture: An Overview. CAST (Council for Agricultural Science and Technology) Issue Paper, No. 23. Faust, M. 2002. New feeds from genetically modified plants: the U.S. approach to safety for animals and the food chain. Livestock Prod. Sci. 74:239-254. FDA. 2006. A Risk-Based Approach to Evaluate Animal Clones and Their Progeny DRAFT. http://www.fda.gov/cvm/CloneRiskAssessment.htm Flachowsky, G., and K. Aulrich. 2001. Nutritional assessment of feeds from genetically modified organism. J. Anim. Feed. Sci. 10:181-194. Flachowsky, G., A. Chesson and K. Aulrich. 2005. Animal nutrition with feeds from genetically modified plants. Arch. Anim. Nutr. 59:1-40. James, C. 2005. Global Status of Commercialized Biotech/GM Crops: 2005. International Service for the Acquisition of Agri-biotech Applications. Brief Number 34-2005. ISAAA, Ithaca, New York. Jaenisch, R., and I. Wilmut. 2001. Developmental biology. Dont clone humans! Science 291:2552. Knowlton, K.F., J.S. Radcliffe, C.L. Novak and D.A. Emmerson. 2004. Animal management to reduce phosphorus losses to the environment. J. Anim. Sci. 82(E. Suppl.):E173-E195. Kuhholzer, B., and R.S. Prather. 2000. Advances in livestock nuclear transfer. Proc Soc Exp Biol Med. 224:240-245. Lei, X.G., and J.M. Porres. 2003. Phytase enzymology, applications, and biotechnology. Biotechol. Lett. 25:1787-1794.

39

Lillico, S.G., A. Sherman, M.J. McGrew, C.D. Robertson, J. Smith, C. Haslam, P. Barnard, P.A. Radcliffe, K.A. Mitrophanous, E.A. Elliot and H.M. Sang. 2007. Oviduct-specific expression of two therapeutic proteins in transgenic hens. Proc. Nat. Acad. Sci. USA: In press. Munkvold, G.P., R.L. Hellmich, and W.B. Showers. 1999. Reduced fusarium ear rot and symptomless infection in kernels of maize genetically engineered for European corn borer resistance. Phytopathol. 7:1071-1077. National Research Council (NRC). 1994. Metabolic Modifiers: Effects on the Nutrient Requirements of Food-producing Animals. The National Academy Press, Washington, D.C. Polejaeva, I.A., S.H. Chen, T.D. Vaught, R.L. Page, J. Mullins, S. Ball, Y. Dai, J. Boone, S. Walker, D.L. Ayares, A. Colman and K.H. Campbell. 2000. Cloned pigs produced by nuclear transfer from adult somatic cells. Nature 407:86-90. The National Academies. 2004. Safety of Genetically Engineered Foods. Approaches to Assessing Unintended Health Effects. Institute of Medicine and National Research Council of The National Academies. The National Academy Press, Washington, D.C. Westhusin. M.E., C.R. Long, T. Shin, J.R. Hill, C.R. Looney, J.H. Pryor and J.A. Piedrahita. 2001. Cloning to reproduce desired genotypes. Theriogenology 55:35-49. Wilmut I., A.E. Schnieke, J. McWhir, A.J. Kind and K.H. Campbell. 1997. Viable offspring derived from fetal and adult mammalian cells. Nature 385:810-813. Willadsen, S.M. 1989. Cloning of sheep and cow embryos. Genome 31:956-962.

40

THE ROLE OF ANIMAL PRODUCTS IN THE DIET TO REDUCE THE RISK OF CHRONIC DISEASE: A FUTURISTIC VISION OF POTENTIAL NEW FOODS Penny M. Kris-Etherton, Deborah Bagshaw, Amy Cifelli Department of Nutritional Sciences The Pennsylvania State University S-126 Henderson Bldg. University Park, PA 16802 Phone: 814-863-2923 Fax: 814-863-6103 Email: pmk3@psu.edu Summary Current dietary guidelines issued for the American public recommend a diet high in fruits, vegetables, whole grains, skim milk/low-fat dairy products, lean meats, and liquid vegetable oils. Animal products can be an important part of a healthy diet and there are many health benefits associated with their consumption. Dairy products have been shown to play a role in reducing blood pressure, controlling body weight and adiposity, improving insulin sensitivity, and reducing risk of colon cancer. Lean meats can increase the protein content of the diet and, thereby, improve plasma lipid and lipoprotein profiles, satiety, and lean body mass. Eggs are an important source of lutein and zeaxanthin, and may decrease risk of age-related macular degeneration. Efforts are ongoing to optimize animal products in terms of their nutrient profiles, and to do this in a way that further reduces the risk of chronic disease. Since animal foods are popular in the U.S. diet, there is great potential for developing designer foods that deliver healthy nutrient profiles. This will be achieved by increasing desired nutrients and reducing/eliminating others (i.e., negative nutrients such as saturated fat and cholesterol). There are different strategies available and being developed that will facilitate achieving even more nutritious animal products. The key will be to develop and implement population-based education programs and campaigns to encourage widespread consumption of these foods, as well as animal products currently in the marketplace, within the context of a healthy diet. Introduction Healthy diet and lifestyle practices are the cornerstones of guidance issued by many organizations for maintaining health and preventing chronic diseases. Guidance has evolved over the years from recommendations that first were based on nutrient profiles to recommendations more recently that have been issued as food-based dietary patterns. Food-based dietary recommendations are easier for consumers to understand and implement than nutrient-specific guidance. The Dietary Guidelines for Americans 2005 included the USDA Food Guide with suggested amounts of food to consume from the basic food groups, subgroups, and oils to meet recommended nutrient intakes at 12 different calorie levels (USDA/HHS, 2005). These recommendations are based on a growing scientific database that has shown many health benefits of a dietary pattern that includes fruits and vegetables, whole grains, low-fat dairy products, lean meats and legumes, and liquid vegetable oils. Animal products are an important aspect of a dietary pattern for health because they are nutrient dense foods that facilitate achieving nutrient adequacy. These foods include lean meats, eggs, and lowfat/non-fat dairy products. Many of the nutrients they provide are important also for decreasing the risk of chronic diseases. The purpose of this paper is to show the role that animal products play in achieving a nutritionally adequate diet and to review the evidence of the health benefits of lean meats, eggs, and lowfat/skim milk dairy products in a dietary pattern that meets current recommendations. This will provide the basis for a discussion about what future designer foods can be developed using animal products that
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

41

facilitate continued evolution of dietary guidance for health. The objective is for future food-based dietary guidance to achieve reductions in chronic disease risk that at the present time are not attainable using a food-based approach. Dietary Guidelines for Americans 2005 Dietary Guidelines for Americans, issued every five years, provide science-based advice to promote health and reduce risk for major chronic diseases through diet and physical activity. Food-based guidance has been issued to promote adherence to a nutritionally adequate diet that meets dietary recommendations to decrease risk of chronic diseases. A basic premise of the Dietary Guidelines is that nutrient needs should be met mainly by consuming foods that comprise a healthful dietary pattern. Two dietary patterns that meet the Dietary Guidelines recommendations and promote healthier lifestyles are the USDA Food Guide and the Dietary Approaches to Stop Hypertension (DASH) Eating Plan (USDA/HHS, 2005). The key recommendations issued in the 2005 Dietary Guidelines are grouped under the following 9 focus areas: Adequate Nutrients within Calorie Needs Consume a variety of nutrient-dense foods and beverages within and among the basic food groups while choosing foods that limit intake of saturated and trans fats, cholesterol, added sugars, salt, and alcohol. Meet recommended intakes within energy needs by adopting a balanced eating pattern such as the USDA Food Guide or the DASH Eating Plan. Weight Management Maintain body weight within a healthy range; balance calories from foods and beverages with calories expended. Prevent gradual weight gain over time; make small decreases in food and beverage calories and increase physical activity. Physical Activity Engage in regular physical activity and reduce sedentary activities to promote health, psychological well-being, and a healthy body weight. Achieve physical fitness by including cardiovascular conditioning, stretching exercises for flexibility and resistance exercise or calisthenics for muscle strength and endurance. Food Groups to Encourage Consume a sufficient amount of fruits and vegetables while staying within energy needs. Choose a variety of fruits and vegetables each day. Consume three or more ounce-equivalents of whole-grain products per day Consume three cups per day of fat-free or low-fat milk or equivalent milk products. Fats Consume less than 10% of calories from saturated fatty acids and less than 300 mg/day of cholesterol, and keep trans fatty acid consumption as low as possible. Keep total fat intake between 20 to 35% of calories with most fats coming from sources of polyunsaturated and monounsaturated fatty acids such as fish, nuts, and vegetable oils. When selecting and preparing meat, poultry, dry beans, and milk or milk products, make choices that are lean, low-fat or fat-free. Limit intake of fats and oils high in saturated and/or trans fatty acids; choose products low in such fats and oils.

42

Carbohydrates Choose fiber-rich fruits, vegetables, and whole grains often. Choose and prepare foods and beverages with little added sugars or caloric sweeteners such as amounts suggested by the USDA Food Guide and the DASH Eating Plan. Reduce the incidence of dental caries by practicing good oral hygiene and consuming sugar- and starch-containing foods and beverages less frequently. Sodium and Potassium Consume less than 2,300 mg (approximately 1 tsp of salt) of sodium per day. Choose and prepare foods with little salt. At the same time, consume potassium-rich foods such as fruits and vegetables. Alcoholic Beverages Those who choose to drink alcoholic beverages should do so sensibly and in moderation defined as the consumption of up to 1 drink per day for women and up to two drinks per day for men. Alcoholic beverages should not be consumed by some individuals, including those who cannot restrict their alcohol intake, women of child bearing age who may become pregnant, pregnant and lactating women, children and adolescents, individuals taking medications that can interact with alcohol, and those with specific medical conditions. Alcoholic beverages should be avoided by individuals engaging in activities that require attention, skill, or coordination, such as driving or operating machinery. Food Safety To avoid microbial foodborne illness: Clean hands, food contacts surfaces and fruits and vegetables. Meat and poultry should not be washed or rinsed. Separate raw, cooked, and ready-to-eat foods while shopping, preparing, or storing foods. Cook foods to a safe temperature to kill microorganisms. Chill (refrigerate) perishable food promptly and defrost foods properly. Avoid raw (unpasteurized) milk or any products made from unpasteurized milk, raw or partially raw eggs or foods containing raw eggs, raw or undercooked meat and poultry, unpasteurized juices and raw sprouts.

As noted, the food-based dietary recommendations are made to achieve nutrient adequacy. Nutrients that are of concern in the U.S. diet for the following groups are: Adults: calcium, potassium, fiber, magnesium, and vitamins A (as carotenoids), C, and E. Children and adolescents: calcium, potassium, fiber, magnesium and vitamin E. Specific population groups including people over age 50, women of childbearing age who may become pregnant as well as those in the first trimester of pregnancy, people with dark skin, and people exposed to insufficient sunlight: vitamin B12, iron, folic acid, and vitamins E and D.

As shown in Tables 1 and 2, animal products including dairy products, eggs, and meats are important sources of many nutrients (USDA, 2005). Both provide good quality protein. Dairy products are the major source of dietary calcium (more than 70%) and also contribute substantial amounts of other nutrients to the diet including: phosphorus (32%), riboflavin (26%), vitamin B12 (21%), protein (19%), potassium (19%), zinc (16%), magnesium (16%), and vitamin A (15%). Fortified milk and ready-to-eat cereals are the predominant food sources of vitamin D. Milk product consumption has been associated with improved diet quality and adequacy of intake of many nutrients, including calcium, potassium, magnesium, zinc, iron, riboflavin, vitamin A, folate, and vitamin D (Huth et al., 2006). Meats are the number one source of protein, zinc, and vitamin B12, and among the top four sources of selenium, iron, 43

vitamin B6, phosphorus, niacin, potassium, and riboflavin (Cotton et al., 2004). Eggs are an excellent source of many essential nutrients, including protein (amino acids), vitamins (A, E, folate, riboflavin, B6, and B12), and minerals (calcium, phosphorus, iron, and zinc) (Korver et al., 2002). As is apparent, animal foods are an important source of many nutrients in the diet. Given the many shortfall nutrients in the U.S. diet (Table 3), and that animal products are important sources of several of these (i.e., calcium, magnesium, folate, vitamin A), strategies are needed to increase consumption of these food sources to improve the nutritional adequacy of the diet. Looking to the future, a question is: can animal products be fortified to provide more shortfall nutrients (Table 3) that typically are not present or are low in meat, dairy products, and eggs? In addition, are there strategies to increase the content of nutrients with demonstrated effects on chronic disease risk reduction?
B

Table 1. Energy, Macronutrients and Selected Micronutrients in Lower Fat, Skim and Full Fat Milk, Cheese, Yogurt and Ice Cream1.
Energy (kcal) Milk, skim, 8 oz. Milk, 1%, 8 oz. Milk, 2%, 8 oz. Milk, whole, 8 oz. American Cheese, low-fat, 1 oz American Cheese, full fat, 1 oz Yogurt, van. low fat, 6 oz Yogurt, low fat w/fruit, 6 oz Yogurt, nonfat, w/fruit, 6 oz Ice cream, 0.5 cup Ice cream, low fat, 0.5 cup
1

Prot (g) 8.2

TF (g) 0.2

CHO (g) 12.1

SFA (g) 0.3

MUFA (g) 0.1

PUFA (g) 0.02

Chol (mg) 5

Ca (mg) 306

Mg (mg) 27

K (mg) 382

P (mg) 247

Na (mg) 103

Sugars (g) 12.5

83

102 122 146

8.2 8.0 7.9

2.4 4.8 7.9

12.2 11.4 11.0

1.5 3.1 4.5

0.7 1.4 2.0

0.10 0.20 0.50

12 20 24

290 285 276

27 27 24

366 366 349

232 229 222

107 100 98

12.7 12.3 12.8

50

6.9

2.0

1.0

1.2

0.6

0.10

10

191

50

231

399

0.2

106

6.3

8.9

0.5

5.6

2.5

0.30

27

175

46

126

184

0.1

288

16.7

4.2

46.8

2.7

1.2

0.10

17

580

54

742

458

224

46.8

356

16.5

4.8

63.0

3.1

1.3

0.10

20

573

54

732

451

220

No data

160

7.5

0.3

32.3

0.2

0.1

0.00

258

26

330

202

99

32.3

266

3.8

17.3

23.8

11.1

4.8

0.70

98

125

12

168

112

65

22.1

125

3.6

3.7

19.6

2.2

1.0

0.20

21

122

11

158

78

56

16.8

kcal=calories, Prot=protein, CHO=carbohydrates, TF=total fat, SFA=saturated fatty acids, MUFA=monounsaturated fatty acids, PUFA =polyunsaturated fatty acids, chol=cholesterol, Ca=calcium, Mg=magnesium, K=potassium, P=phosphorus, Na=sodium.

44

Table 2. Energy, Macronutrients and Selected Micronutrients in Eggs and Popular Cuts of Meat1
Energy (kcal) Hamburger, 30% fat, 3 oz Hamburger, 15% fat, 3 oz. Hamburger, 5% fat, 3 oz Beef, top sirloin, lean, 3 oz Beef, eye round, lean, 3 oz Chicken Breast, skinless, 3 oz Chicken Thigh, skinless, 3 oz Pork chop, lean, 3 oz Lamb chop, lean, 3 oz Egg, large, 50 g2
1

Prot (g) 21.6 22.0

TF (g) 15.5 13.1

SFA (g) 6.2 5.0

MUFA (g) 7.5 5.7

PUFA (g) 0.4 0.4

Chol (mg) 70 77

Fe (mg) 1.91 2.21

Zn (mg) 5.2 5.4

Thiamin (mg) 0.04 0.04

Ribofl (mg) 0.15 0.15

Niacin (mg) 3.87 4.57

B12 (mcg) 2.46 2.24

232 213

145 180

22.3 24.9

5.6 8.2

2.5 3.2

2.3 3.4

0.3 0.3

65 62

2.41 1.61

5.5 4.5

0.04 0.07

0.15 0.12

5.05 6.67

2.10 1.47

143

24.8

4.1

1.5

1.7

0.1

46

2.01

4.1

0.06

0.13

4.33

1.38

142

26.7

3.1

0.9

1.1

0.7

73

0.89

0. 9

0.06

0.98

11.79

0.29

174

21.6

9.0

2.5

3.4

2.1

79

1.10

2.1

0.06

0.19

5.43

0.26

199 160 75

21.6 23.5 6.3

11.8 6.6 5.0

4.3 2.7 1.6

5.1 2.7 1.9

0.9 0.3 0.7

71 72 213

0.93 2.10 0.72

3.4 4.4 0.6

0.63 0.11 0.03

0.30 0.36 0.25

3.87 4.90 0.04

0.80 2.70 0.50

kcal=calories, Prot=protein, TF=total fat, SFA=saturated fatty acids, MUFA=monounsaturated fatty acids, PUFA =polyunsaturated fatty acids, chol=cholesterol, Fe=iron, Zn=zinc, Ribofl=riboflavin, B12=Vitamin B12. 2 From Kerver et al. (2002).

45

Table 3. Probabilities of Adequacy for Selected Nutrients on the First 24-hour Recall among Adult CSFII 1994-96 Participants1,2 Probability of adequacy (%) Nutrient Men Women Vitamin A Vitamin C Vitamin E Thiamin Riboflavin Niacin Folate1 Vitamin B-6 Vitamin B-12 Phosphorus Magnesium Iron Copper Zinc Calcium
1 2

47.0 49.3 14.1 83.9 85.8 90.5 33.9 78.3 80.5 94.3 36.1 95.5 87.4 65.7 58.6

48.1 52.3 6.8 72.2 80.9 80.4 20.9 60.7 64.2 85.1 34.3 79.4 73.3 62.0 45.7

From: Foote et.al. (2004). Nutrients considered shortfall nutrients (Dietary Guidelines Advisory Committee, 2005) in bold.

Health Benefits of Dairy Products, Meats/Meat Products, and Eggs Dairy Products Many health benefits are associated with consumption of dairy products (Huth et al., 2006). These are related to the nutrients and bioactive lipids and protein components in dairy products. In addition, consumption of dairy products displaces other foods, such as sugar-sweetened carbonated beverages and, therefore, results in health benefits. Epidemiologic, clinical trials, animal studies, and in vitro experiments show that consumption of/constituents in dairy products help reduce the risk of chronic diseases including osteoporosis, hypertension, excess body weight and fat, insulin resistance syndrome, and some cancers (Huth et al., 2006). Osteoporosis In a 2004 report on bone health and osteoporosis (U.S. Department of HHS, 2004), it is estimated that in 2020 one in two Americans over the age of 50 will have, or be at high risk of developing osteoporosis. Currently, osteoporosis is responsible for approximately 1.5 million spontaneous bone fractures yearly with an estimated annual health care impact of $17 billion. As noted in this report, much can be done with diet and lifestyle practices to decrease risk, and consequently prevalence of osteoporosis. Notably, attaining an optimal peak bone mass for maximal bone strength later in life for greater resistance to fracture is at the forefront of intervention strategies. Dairy products play a key role in the production and maintenance of a healthy bone matrix throughout life. Huth et al. (2006) reported a positive relationship between calcium intake (as dairy products in many studies) and bone health in 68 of 70 controlled intervention studies. This effect not only reflects calcium intake but also vitamin D, potassium and magnesium, as well as high quality protein, all essential nutrients for optimal bone health. Hypertension Hypertension is a major health problem in the U.S., affecting about 29% of the population (CDC, 2005). Randomized clinical trials show that diets high in calcium or dairy products decrease blood pressure. The DASH Study, a landmark clinical trial, demonstrated that a diet that emphasized fruits, vegetables, and low-fat dairy products and included whole grains, poultry, fish, and nuts had a remarkable blood pressure-lowering effect (similar to that typically 46

seen with pharmacologic therapy) (Appel et al., 1997; Appel, 2000). Among all participants, the DASH diet significantly lowered mean systolic blood pressure by 5.5 mmHg and mean diastolic blood pressure by 3.0 mmHg (net of control) (Appel et al., 1997). Body Weight/Body Composition Population studies consistently demonstrate a beneficial association between calcium intake, particularly from dairy foods, and lower body weight and lower body fat (Huth et al., 2006). In the Coronary Artery Risk Development in Young Adults (CARDIA) Study (Pereira et al., 2002), 3,000 adults (ages 18-30) were followed for 10 years. The risk of weight gain was 67% lower in persons who consumed the most dairy foods versus those who consumed the least. Clinical trial evidence from Zemel et al. (2004) indicated that a high calcium (1,200-1,300 mg/day) weight loss diet and high dairy (3-4 servings of dairy foods/day) weight loss diet resulted in significantly greater weight and body fat loss than a low calcium (400-500 mg/day) weight loss diet. The high dairy diet resulted in the greatest weight loss. All diets were reduced by 500 calories/day for 24 weeks to achieve weight loss. The low calcium diet group lost 14.5 lbs (body fat lost was 10.6 lbs) versus the high calcium diet group that lost 18.9 lbs body weight (12.3 lbs body fat) and the high dairy group that lost 24.4 lbs (15.8 lbs body fat). Subjects on the low calcium diet lost 5.3% of their visceral fat, while those consuming the high calcium and high dairy diets lost 12.9% and 14.0%, respectively, of their visceral fat. This has important implications because abdominal obesity is a major risk for insulin resistance syndrome (IRS). Insulin Resistance Syndrome IRS, also known as metabolic syndrome, is a condition that is associated with a number of risk factors, including obesity, insulin resistance, elevated plasma insulin levels, low blood HDL-cholesterol and high circulating triglyceride levels, hypertension, and impaired fibrinolytic capacity (Reaven, 1993). Individuals with IRS are at high risk for developing diabetes and heart disease. In the CARDIA Study, increased dairy consumption was inversely associated with IRS among overweight adults (Pereira et al., 2002). Each additional serving of dairy products was associated with a 21% lower odds ratio of having IRS. In addition, Liu et al. (2006) recently showed a 21% decreased risk of type 2 diabetes in women in the highest quintile of dairy product intake versus those in the lowest quintile of intake. Each daily serving increase in dairy intake was associated with a 4% lower risk of type 2 diabetes (Liu et al., 2006). Cancer An inverse association has been reported between the intake of dairy products and colorectal cancer (Alvarez-Leon et al., 2006). Slattery et al. (1997) reported an inverse association between dietary calcium (> 800 mg/day) from milk and other dairy products and risk of colon cancer, especially in males. A protective effect against colon cancer was observed for yogurt that was independent of its calcium content in a large case-control study conducted in California reviewed by Huth et al. (2006). Clinical trial evidence also shows a beneficial effect of supplemental calcium and dairy foods on colonic epithelial cell proliferation and a restoration of normal cell differentiation in patients with a history of developing intestinal polyps or noncancerous growths in the colon (Holt et al., 2001). There is some evidence of an association between high-fat dairy products and risk of prostate cancer. In a prospective study in France, a higher risk of prostate cancer was observed among subjects with higher dairy product consumption (Kesse et al., 2006). In addition, a harmful effect of yogurt consumption on prostate cancer also was reported. A study in the U.S. reported that very high intakes of dietary calcium (> 2 g/day) as well as low intakes (< 700 mg/d) were associated with increased risk of prostate cancer (Rodriguez et al., 2003). Of importance is that moderate levels of dietary calcium were not associated with increased risk of prostate cancer. In addition, dairy intake was not associated with prostate cancer risk (Rodriguez et al., 2003). A recent large (10,000 subjects) prospective study reported that neither increased dairy food intake nor increased calcium intake from supplements was associated with prostate cancer risk (Koh et al., 2006). There is no evidence of an association between the consumption of dairy products and breast cancer.

47

Meats/Meat Products
Meat is a good source of protein and micronutrients (Table 2). Beef also is the number one source of saturated fat in the diet. Collectively, meat and dairy products contribute 60% of the saturated fat in the U.S. diet. Efforts to decrease saturated fatty acids target red meat reduction (as well as full-fat dairy products). There is evidence that lean meat products, however, beneficially affect lipids and lipoproteins as part of a blood cholesterol-lowering diet (see below). In addition to beneficially affecting heart disease risk factors, a diet higher in protein comprised of lean meats favorably affects body composition during weight maintenance as well as weight loss. For the control of body weight, meat and meat products may beneficially affect satiety. Studies have been conducted to evaluate the role of meats and meat products on cancer risk (see below). Cardiovascular disease In the Nurses Health Study, high protein intakes were associated with a low risk of ischemic heart disease (Hu et al., 1999). When extreme quintiles of total protein intake were compared, the relative risk was 0.74 indicating that risk was decreased by 26%. In a controlled clinical trial that evaluated diets high in carbohydrate, protein (half from plant sources) and unsaturated fat (mainly monounsaturated fat), substituting protein for carbohydrate resulted in decrease blood pressure and serum levels of LDL-cholesterol, HDL-cholesterol, and triglycerides (Appel et al., 2005). The authors concluded that partial substitution of carbohydrate with either protein (or monounsaturated fat) can reduce estimated CVD risk due to favorable effects on multiple cardiovascular disease risk factors. Clinical studies also have been conducted to evaluate inclusion of lean red meat versus lean white meat on serum lipids and lipoproteins in individuals with high blood cholesterol levels because health professionals frequently counsel these patients to avoid red meat. Studies to date have shown similar hypocholesterolemic effects of recommended diets that contain either lean red meat or lean white meat (Scott et al., 1994; Davidson et al., 1999; Hunninghake et al., 2000). These studies thus showed that lean meat can be a part of a heart healthy diet. The important point is that lean meat be included since saturated fat is high in some cuts of higher-fat red meat. Weight Loss, Body Composition and Satiety Increased dietary protein may favorably affect body weight regulation through effects on satiety, thermogenesis and substrate partitioning (i.e., favor fat oxidation rather than deposition in adipose tissue) (Blaak, 2006). A meta-analysis conducted that evaluated the effects of low carbohydrate (higher protein) versus low fat weightloss diets reported a greater weight loss after 6 months (- 3.3 kg) on the higher protein diets (Nordmann et al., 2006). However, the difference was no longer apparent after 12 months on the low carbohydrate versus low fat weight loss diets. In a study by Volek et al. (2004), both men and women on a very low carbohydrate ketogenic weight loss diet lost more trunk fat than did subjects on a low fat weight loss diet. Layman et al. (2005) reported that subjects on a high protein, reduced carbohydrate diet lost more fat mass and tended to lose less lean body mass than did subjects on the low fat (high carbohydrate) weight loss diet. Cancer There is epidemiologic evidence that consumption of processed meats and red meat increase risk of colorectal cancer (Giovannucci, 2003). Questions remain about whether processing per se, and cooking techniques (i.e., that result in a smoked/charred surface), are a factor that explains the epidemiologic evidence. A recent study reported that high red meat intake (> 1.5 servings per day) increased risk about 2-fold of estrogen and progesterone receptor positive breast cancer (Cho et al., 2006).

48

Eggs In addition to serving as a dietary source of many essential nutrients (Table 2), eggs can be enriched with omega-3 fatty acids, both plant-derived and marine-derived, through modification of the laying hens diet. Typically, flaxseed is used to increase -linolenic acid (ALA), and fish meal/oil is used to increase long-chain omega-3 fatty acids in egg yolk. Long-chain omega-3 fatty acids have been shown to decrease risk of cardiovascular disease (Kris-Etherton et al., 2002). Although ALA can be converted to long chain omega-3 fatty acids in vivo, this process is not efficient in either the laying hen or man. Egg yolk is a highly bioavailable source of lutein and zeaxanthin, two dietary carotenoids that accumulate in the macula. Research has shown that increased blood levels of lutein and zeaxanthin might decrease the risk of age-related macular degeneration, the leading cause of blindness in the Western world. A recent study showed that blood levels and macular pigment optical density (a measure of carotenoid levels in the macula) were improved significantly in women whose diets were supplemented with six carotenoid-enriched eggs per week (Wenzel et al., 2006). This is important because lutein supplementation may be effective in slowing the progression of age-related macular degeneration. However, in many of the recent clinical trials with carotenoid-enriched eggs, participants who were classified as hyperresponders experienced marked elevations in plasma total- and/or LDLcholesterol levels. An example of the heterogeneity of response to dietary cholesterol in a healthy population of 40 men and 51 women has been reported recently (Herron et al., 2006). In addition to carotenoids and omega-3 fatty acids, it is possible to manipulate the levels of many other nutrients in eggs. Moreover, in the future, eggs will most likely be enriched with other substances that enhance human health. A Look to the Future An ongoing goal in the nutrition field an ongoing goal is to improve dietary recommendations with the intent to achieve optimal health and prevent chronic disease. The evidence is clear that a nutrientrich diet can have a huge impact on health. At the present time, we have a good understanding of what a healthy dietary pattern is. The challenge is to improve adherence to current dietary recommendations. Looking to the future and realizing that meats, dairy products and eggs are popular foods, efforts are ongoing to modify nutrient profile of these foods. This is being done in a variety of ways including animal feeding and breeding practices, genetic modification of plant and animal foods, and a variety of practices that manipulate nutrient profile such as reducing/deleting fat and cholesterol and increasing omega-3 fatty acids. A key question is what the cumulative impact of nutrient changes in these foods will have on the overall diet and, in turn, its impact on health. The challenge is not the evolution of science but, rather, educating consumers about new designer foods and how they can be incorporated into a healthy diet. References Alvarez-Leon, E.E., B. Roman-Vinas and L. Serra-Majem. 2006. Dairy products and health: A review of the epidemiological evidence. Br. J. Nutr. 96 Suppl: S94-99. Appel, L.J. 2000. The role of diet in the prevention and treatment of hypertension. Curr. Atheroschler. Rep. 2:521-528. Appel, L.J., T.J. Moore, E. Obarzanek, W.M. Vollmer, L.P. Svetkey, F.M. Sacks, G.A. Bray, T.M. Vogt, J.A. Cutler, M.M. Windhauser, P.H. Lin and N. Karanja. 1997. A clinical trial of the effects of dietary patterns on blood pressure. DASH Collaborative Research Group. N. Engl. J. Med. 336:1117-1124. Appel, L.J., F.M. Sacks, V.J. Carey, E. Obarzanek, J.F. Swain, E.R. Miller, 3rd, P.R. Conlin, T.P. Erlinger, B.A. Rosner, N.M. Laranjo, J. Charleston, P. McCarron and L.M. Bishop. 2005. Effects of protein, monounsaturated fat, and carbohydrate intake on blood pressure and serum lipids: results of the omniheart randomized trial. JAMA. 294:2455-2464.

49

Blaak, E.E. 2006. Prevention and treatment of obesity and related complications: a role for protein? Int. J. Obes. (Lond). 30 Suppl:S24-27. CDC. 2005. Racial/Ethnic disparities in prevalence, treatment, and control of hypertension --- United States, 1999-2002. MMWR Weekly. 54:7-9. Cho, E., W.Y. Chen, D.C. Hunter, M.J. Stampfer, G.A. Colditz, S.E. Hankinson and W.C. Willett. 2006. Red meat intake and risk of breast cancer among premenopausal women. Arch. Intern. Med. 166:2253-2259. Cotton, P.A., A.F. Subar, J.E. Friday and A. Cook. 2004. Dietary sources of nutrients among US adults, 1994 to 1996. J. Amer. Diet. Assoc. 104:921-930. Davidson, M.H., D. Hunninghake, K.C. Maki, P.O. Kwiterovich, Jr., and S. Kafonek. 1999. Comparison of the effects of lean red meat vs lean white meat on serum lipid levels among free-living persons with hypercholesterolemia: a long-term, randomized clinical trial. Arch. Intern. Med. 159:1331-1338. Foote, J.A., S P. Murphy, L.R. Wilkens, P.P. Basiotis and A. Carlson. 2004. Dietary variety increases the probability of nutrient adequacy among adults. J. Nutr. 134:1779-1785. Giovannucci, E. 2003. Diet, body weight, and colorectal cancer: A summary of the epidemiological evidence. J. Womens Health (Larchmt). 12:173-182. Herron, K.L., I.E. Lofgren, X. Adiconis, J.M. Ordovas and M.L. Fernandez. 2006. Associations between plasma lipid parameters and APOC3 and APOA4 genotypes in a healthy population are independent of dietary cholesterol intake. Atherosclerosis 184:113-120 Holt, P.R., C. Wolper, S.F. Moss, K. Tang and M. Lipkin. 2001. Comparison of calcium supplementation or low-fat dairy foods on epithelial cell proliferation and differentiation. Nutr. Cancer. 41:150-155. Hu, F.B., M.J. Stampfer, J.E. Manson, E. Rimm, G.A. Colditz, F.E. Speizer, C.H. Hennekens and W.C. Willett. 1999. Dietary protein and risk of ischemic heart disease in women. Am. J. Clin. Nutr. 70:221227. Hunninghake, D.B., K.C. Maki, P.O. Kwiterovich, Jr., M.H. Davidson, M.R. Dicklin and S.D. Kafonek. 2000. Incorporation of lean red meat into a National Cholesterol Education Program Step I diet: a long-term, randomized clinical trial in free-living persons with hypercholesterolemia. J. Am. Coll. Nutr. 19:351-360. Huth, P.J., D.B. DiRienzo and G.D. Miller. 2006. Major scientific advances with dairy foods in nutrition and health. J. Dairy Sci. 84:1207-1221. Kerver, J.M., Y. Park and W.O. Song. 2002. The role of eggs in American diets: Health implications and benefits. Pages 9-18. In: Eggs and Health Promotion. Watson, R.R. (ed.). Iowa State Press, Ames, IA. Kesse, E., S. Bertais, P. Astorg, A. Jaouen, N. Arnault, P. Galan and S. Hercberg. 2006. Dairy products, calcium, and phosphorus intake, and the risk of prostate cancer: Results of the French prospective SU.VI.MAX (Supplementation en Vitamines et Mineraux Antioxydants) study. Br. J. Nutr. 95:539545. Koh, K.A., H.D. Sesso, R.S. Paffenbarger, Jr., and I.M. Lee. 2006. Dairy products, calcium and prostate cancer risk. Br. J. Cancer. 95:1582-1585. Kris-Etherton, P.M., W.S. Harris and L.J. Appel. 2002. Fish consumption, fish oil, omega-3 fatty acids, and cardiovascular disease. Circulation. 106:2747-2757. Layman, D.K., E. Evans, J.I. Baum, J. Seyler, D.J. Erickson and R.A. Boileau. 2005. Dietary protein and exercise have additive effects on body composition during weight loss in adult women. J. Nutr. 135:1903-1910. Liu, S., H.K. Choi, E. Ford, Y. Song, A. Klevak, J.E. Buring and J.E. Manson. 2006. A prospective study of dairy intake and the risk of type 2 diabetes in women. Diabetes Care. 29:1579-1584. Nordmann, A.J., A. Nordmann, M. Briel, U. Keller, W.S. Yancy, Jr., B.J. Brehm and H.C. Bucher. 2006. Effects of low-carbohydrate vs low-fat diets on weight loss and cardiovascular risk factors: A metaanalysis of randomized controlled trials. Arch. Intern. Med. 166:285-293.

50

Pereira, M.A., D.R. Jacobs, Jr., L. Van Horn, M.L. Slattery, A.I. Kartashov and D.S. Ludwig. 2002. Dairy consumption, obesity, and the insulin resistance syndrome in young adults: The CARDIA study. JAMA. 287:2081-2089. Reaven, G.M. 1993. Role of insulin resistance syndrome in human disease (syndrome X): An expanded definition. Ann. Rev. Med. 44:121-131. Rodriguez, C., M.L. McCullough, A.M. Mondul, E.J. Jacobs, D. Fakhrabadi-Shokoohi, E.L. Giovannucci, M.J. Thun and E.E. Calle. 2003. Calcium, dairy products, and risk of prostate cancer in a prospective cohort of United States men. Cancer Epidemiol. Cancer Prev. 12:597-603. Scott, L.W., J.K. Dunn, H.J. Pownall, D.J. Brauchi, M.C. McMann, J.A. Herd, K.B. Harris, J.W. Savell, H.R. Cross and A.M. Gotto, Jr. 1994. Effects of beef and chicken consumption on plasma lipid levels in hypercholesterolemic men. Arch. Intern. Med. 154:1261-1267. Slattery, M.L., B.J. Caan, D. Potter, T.D. Berry, A. Coates, D. Duncan and S.L. Edwards. 1997. Dietary energy sources and colon cancer risk. Am. J. Epidemiol. 145:199-210. USDA. 2005. National Nutrient Database. http://www.nal.usda.gov/fnic/foodcomp/search/. USDA/HHS. 2005. Dietary Guidelines for Americans. http://www.healthierus.gov/. USDHHS. 2004. Bone Health and Osteoporosis: A Report of the Surgeon General. Volek, J., M. Sharman, A. Gomez, D. Judelson, M. Rubin, G. Watson, B. Sokmen, R. Silvestre, D. French and W. Kraemer. 2004. Comparison of energy-restricted very low-carbohydrate and low-fat diets on weight loss and body composition in overweight men and women. Nutr. Metab. (Lond). 1:13. Wenzel, A.J., C. Gerweck, D. Barbato, R.J. Nicolosi, G.J. Handelman and J. Curran-Celentano. 2006. A 12-wk egg intervention increases serum zeaxanthin and macular pigment optical density in women. J. Nutr. 136:2568-2573. Zemel, M.B., W. Thompson, A. Milstead, K. Morris and P. Campbell. 2004. Calcium and dairy acceleration of weight and fat loss during energy restriction in obese adults. Obes. Res. 12:582-590.

51

WHAT DO WE KNOW AFTER 30 YEARS OF EARLY NUTRITION RESEARCH?


J. J. Dibner Novus International, Inc. 20 Research Park Dr. Missouri Research Park St. Charles, MO 63304 Phone: 636-926-7410 FAX: 636-926-7405 Email: julia.dibner@novusint.com Summary Studies of the effect of early nutrition on development and function of the digestive and immune systems have yielded enormous advances in our understanding of the role of enteral intake, specific nutrients, and even ingested bacteria on animal growth and resilience. This review will first briefly cover the status of the gastrointestinal (GI) system at hatch and effects of early nutrition, including in ovo feeding, on its development. Second, we will consider the importance of delayed access to feed on subsequent growth and efficiency. Finally, the interaction between early intake and the accompanying exposure to bacteria and their products will be discussed. Nutritional status of the neonate can influence the composition of gastrointestinal microflora that subsequently colonizes the gut. This in turn influences the development of immune sensitization to or tolerance of orally ingested antigens, usually for the rest of the animals life. Clearly, influencing the heterogeneity and stability of that microflora is a critical aspect of early nutrition because it can greatly impact future animal health, growth, and efficiency. Introduction Poultry have been selected for more than 50 years for muscle accretion. Along with this positive selection pressure for meat production, other physiological systems may not have been under selection per se but have evolved to support the increasing rate of muscle growth. Often these support systems, such as the cardiovascular, immune and skeletal systems, are only marginally able to meet the pace set by muscle growth. As a result, failures in support systems have occurred, for example the development of ascites associated with cardiovascular insufficiency. In response, more recent selection programs often include skeletal, immune or cardiovascular system components. Ultimately, all of these support organs rely on the provision of metabolic precursors by the gastrointestinal system to support their own growth and development. Although not directly monitored in genetic programs, the development of the digestive system has been revealed as crucial to the realization of a birds genetic potential for growth. With increasing growth rates and precocity in meat poultry, the first week of life has become a greater proportion of the total lifespan, often representing 20% of the total. Thus, adequate nutrition during the first days after hatch has become a necessary component of a successful production cycle. This has prompted a renewed interest in the development of the GI system both shortly before, as well as after, hatch. Studies of the nutritional requirements of the hatchling and the consequences of delaying the onset of enteral nutrition have expanded the body of published research in early nutrition of poultry over the past decade. Gastrointestinal Development Hatchling poultry undergo a fundamental change in nutrient supply and metabolism upon emergence from the egg shell. During embryonic incubation, yolk lipid supplies the caloric needs of the bird and is delivered from the yolk sac via the blood stream. As hatch approaches, glycogen accumulates in the liver and is used during the hatching process (Freeman and Vince, 1974). In addition, lipid is transferred to the liver of the embryo and the yolk sac is internalized within the body wall (Romanoff,
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

52

1960). If the animal is denied food and water, this lipid and the residual yolk water and protein can supply nutrients to the hatchling until feed and water become available; however, the residual yolk nutrients are more valuable for their ability to confer passive immunity and as structural material for the developing bird (Dibner et al., 1998b). Evidence confirms that some of the residual yolk makes its way into the intestine via the yolk stalk and, thus, provides digestible nutrients that may stimulate maturation of the digestive and absorptive functions of the neonatal intestine (Noy et al., 1996; Sulaiman et al., 1996). Thus, one possible way to provide nutrients to the neonate, particularly in trace amounts, is through supplementation of the breeder hen diet with nutrients that are transferred to the yolk sac. There has been some work done showing the benefits of increasing availability of trace minerals to the progeny through feeding organic trace minerals to the breeder hen (Kidd et al., 2000). These studies have shown benefits in early innate immune capability in the progeny. There is clearly a need for more work of this kind. Remarkably, the post hatch growth of the broiler and turkey poult small intestine relative to body weight is on the order of 7 fold for the first two weeks of life (Iji et al., 2001; Uni et al., 1999). Accompanying this increase in weight is an increase in functional capacity. The general direction of gut development is anterior to posterior, with the foregut being the most differentiated at the time of hatch (Romanoff, 1960). Although the bird has not ingested feed at the time of hatch, intestinal and pancreatic enzymes as well as nutrient transport capabilities are already present (Buddington, 1992; Sklan et al., 2003). Nutrient digestibility, however, is not fully mature at this time (Krogdahl and Sell, 1989). Ontogenetic changes occur, both before and after hatch, that include increased levels of pancreatic and intestinal enzymes (Noy and Sklan, 1995; Sell et al., 1991; Sklan, 2001)), increases in overall GI tract surface area for absorption (Iji et al., 2001; Nitsan et al., 1991; Sklan and Noy, 2003b), and changes in nutrient transporters (Buddington and Diamond, 1989; Sklan et al., 2003). One avenue to increase the maturity of the digestive system at hatch would be to stimulate its development when the bird is still in the shell. This has been explored by Uni and Ferket in several studies (Ferket, 2006; Uni and Ferket, 2004). Based on the observation that immediate access to feed is necessary to support optimum growth of both the support systems such as the digestive system (Baranyiova, 1972; Baranyiova and Holman, 1976), and the demand organs such as muscle and skeleton (Halevy et al., 2003; Mozdziak et al., 2002a, 2002b; Ohlsson and Smith, 2001), several investigators have examined the effects of providing soluble nutrients and growth factors in ovo (Ferket, 2006; Foye et al., 2006; Peebles et al., 2001; Smirnov et al., 2006; Tako et al., 2004, 2005). The nutrients provided in ovo would be expected to be delivered to the embryo orally in conjunction with its oral consumption of amniotic fluid beginning on day 14 of incubation. A number of markers of enteric development were affected by in ovo feeding of solutions of various carbohydrates and protein (Uni and Ferket, 2004). There was a trophic effect on the gut resulting in increased jejunal villus length compared to control or saline injected controls, as well as enhanced expression of genes for brush border enzymes such as sucrase-isomaltase and leucine aminopeptidase, and for nutrient transporters such as SGLT-1 and PEPT-1 (Foye et al., 2006; Tako et al., 2005). One result of this substrate provision and relative increased maturation was an increase in liver glycogen at hatch (Uni and Ferket, 2004). This could represent a critical reserve for the hatchling in the period between hatch and the availability of feed. There appears to be promise in this approach to increasing GI system maturation at hatch. Nutrient Digestibility in the Neonate One way to evaluate development of the gut is to test performance and nutrient digestibility as a function of age. Batal and Parsons (2002) tested digestibility of a corn soybean meal (SBM) diet in commercial broiler males on days 0-2, 3-4, 7, 14 and 21 days of age. The diet provided 23% crude protein (CP) and 3168 kcal/kg total metabolizable energy (TMEn). They observed that metabolizable energy (MEn) significantly increased with time and reached 3377 kcal/kg dry matter (DM) on day 21. In addition, digestibility of individual amino acids also significantly increased with time, with increases ranging from

53

2 to 7% over the 21 day study. These studies indicate that hatchling birds are not able to use nutrients as efficiently as older birds. Noy and Sklan (2002) approached the question in a different way, feeding corn gluten meal-SBM diets that contained 3, 7 or 11% fat (soybean oil) and 18, 23 or 28% protein. They observed that increasing fat or protein decreased feed intake and also decreased body weight (BW). The optimum body weight and feed intake was observed at 18% CP and 3% fat. Interestingly, the higher protein and fat diets did not change carcass protein and fat percent. The highest protein and fat retentions were also observed in the 18% protein 3% fat fed birds. In a subsequent study, the diets contained 18, 21, 24 or 28% protein and 4.5 or 9% fat. They observed that the highest 7 day BW occurred in the birds fed 21% protein 4.5% fat diets. When they added an inert ingredient (cellulose at 3 or 13%) to the diet they also observed a decrease in feed intake. Unlike older birds, hatchlings appeared to be unable to compensate for a decrease in nutrient density by increasing feed intake. Similar results were obtained in studies of turkey poults (Noy and Sklan, 2004). These results indicate that increased nutrient density, although sometimes accompanied by increased feed or caloric efficiency, can result in a decrease in feed intake, particularly if the fat or protein is simply added on top of a diet formulated to meet NRC recommendations. Role of Specific Nutrients in Hatchling Feed Energy It is well accepted that fat digestibility improves over the first week of life in broiler chicks (Lilburn, 1998). Digestibility is particularly low for more saturated fats compared to vegetable oils. This can be overcome through the addition of bile salts to the diet (Gomez and Polin, 1976). Availability of unsaturated fatty acids is actually much higher than that for intact fats, even those found in relatively saturated fats such as tallow (Noy and Sklan, 1995, 1998, 1999; Sklan, 2003). Turner et al. (1999) compared the digestibility of carbohydrate from corn with supplemental fat from an animal vegetable blend or from a synthetic fat based on medium chain triglycerides (MCT). Overall digestibility of nonlipid DM was higher in the corn based diet than either fat based diet. Interestingly, the carbohydrate based diet gave a higher apparent metabolizable energy than the fat based diets in the 3-5-day time frame. In addition, the MCT-based synthetic fat was significantly more available than the animal vegetable blend during the same early time frame. This agrees well with the improvement of absorption with bile salt supplementation described earlier. Batal and Parsons (2004) investigated the relative use of various carbohydrate sources as affected by age in the chick. They compared dextrose, conventional cornstarch, dextrinized cornstarch, corn-syrup solids, pregelatinized unmodified cornstarch, pregelatinized tapioca starch, tapioca dextrin, high-amylose starch and a mixed glucose polymer (polycose) in a corn-SBM diet over the 0-7 day time frame. MEn for the diet containing dextrose (3313 kcal/kg DM) was significantly greater than that for any other carbohydrate during the 0-2 day period. Cornstarch, dextrinized corn starch, and tapioca dextrins were also above the 3000 kcal/kg DM level, with lowest MEn seen with the high-amylose starch diet. In a subsequent study, performance and MEn yield of SBM diets containing various high-carbohydrate ingredients (dextrose, conventional corn, waxy corn, high-oil corn, corn flour and rice flour) were compared. As in the previous experiment, dextrose based diets gave the highest MEn levels over the first week. Diets made using corn flour were not significantly different at 0-2 days but were significantly lower at 3-4 days. Lower MEn was seen with waxy corn and rice flour. There were no performance differences over the first 7 days but over the 0-3 wk time frame, dextrose, waxy corn and high oil corn gave significantly better weight gain than other carbohydrate sources, including conventional corn. Over the same period, feed efficiency was significantly poorer for corn and rice flour than for the other carbohydrate sources (Batal and Parsons, 2004).

54

Specific Amino Acids Studies of lysine and sulfur amino acid use in corn-SBM diets during the neonatal period were reported by Sklan and Noy (2003a) and by Garcia and Batal (2005). The studies had somewhat different objectives in that Sklan and Noy were looking at increasing levels of crude protein from 16 to 28% of the diet and testing the effect of that when lysine, sulfur amino acids (SAA) and threonine levels were either maintained as recommended by the NRC (g/kg diet) or were kept at a constant ratio to CP (AA:CP). The ratio to CP values used were that found in the NRC for a 23% CP diet and were 9 g/kg SAA and 11 g/kg for lysine, both at 3198 kcal/kg ME (Sklan and Noy, 2003a). Results indicated that performance during the first 7 days could be improved with higher CP, providing the AA:CP was balanced. Performance was increased up to the highest CP level tested, 26%. In another experiment, the AA:CP ratio was kept constant as described above and dietary CP levels of 16, 20, 24, and 28% were fed in diets containing either 2997 or 3198 kcal/kg ME. In this study, birds were carried to 40 days of age to test the retention of any benefit to market age. In fact, the greatest BW was seen with the highest CP, highest ME diet and the effect persisted to 40 days. Initially, the best feed efficiency was also seen in the highest density diet, but by 40 days of age the feed efficiencies were not different. Garcia and Batal (2005) tested the hypothesis that, based on nutrient uptake and digestibility differences that occur with age, the digestible lysine (DLYS) and SAA (DSAA) requirements might be expected to be higher from day 0-7 than from day 0-21 or days 7-21. The authors point out that most of the NRC requirements were based on studies conducted between 7 and 21 days to avoid the complication of variable amino acid contributions from the residual yolk. All corn-corn gluten meal-SBM diets were formulated to contain 23% CP. Results across the 21 day study agreed with the essential amino acid requirements as given in the NRC. Surprisingly, the authors observed no difference in DLYS and DSAA requirements as determined at 4, 7 or 21 days. These results suggest that the increases in digestibility associated with increasing bile salt availability, digestive enzyme activity, and nutrient transport capacity over the first 7 to 10 days post hatch may not impact nutrient requirements or performance at levels that can be detected in growth studies. More work on specific nutrient requirements in the neonate and their potential to impact long term health and performance needs to be done. It is crucial that studies include other systems, both support and demand, in addition to performance efficiency in the criteria for determination of requirements. Some work on glutamine (Yi et al., 2005), minerals (Kidd et al., 2000; Dibner, 2005) and vitamins (McDowell, 2000) has been published but much more work directed at early post-natal life needs to be done. Enteral Nutrition If there is a single most important conclusion in 30 years of early nutrition research, it is the essentiality of oral intake of nutrients by the neonate. When there is a delay in the intake of feed and water, consequences include poor response to vaccination, slow GI and immune development, poor disease resistance, and also poor long-term growth performance in both chicks and poults (Bar-Shira and Friedman, 2005; Dibner and Knight, 2001; Dibner et al., 1998a; Juul-Madsen et al., 2004; Noy and Sklan, 1998). This early enteral feeding cannot be achieved through provision of an inert substance or of water alone, there must be nutrient value in the offering (Noy and Sklan, 1998, 2002). Not enough work has been done to say definitely what the optimum nutrient balance would be, although early work defining the nutritional profile for the high moisture supplement OASIS indicated that best intake and performance in broilers could be achieved with a high carbohydrate, high protein supplement containing only the fat found in the ingredients (Dibner et al., 1998a), and that the presence of organic acids in the Oasis formulation might provide disease resistance benefits related to the establishment of an acid tolerant microflora (Jackson, 2005; Yi et al., 2005).

OASIS is a registered trademark of Novus International and is registered in the United States and other countries.

55

Delayed access to feed has severe effects on the development of the GI and immune systems. Simple reductions in organ weight for both systems have been reported by many researchers (Dibner et al., 1998b; Mikec et al., 2006; Noy et al., 2001; Noy and Sklan, 1998, 1999; Potturi et al., 2005; Sklan and Noy, 2000). Noy et al., (2001) found that delayed provision of nutrients for 48 hr affected the growth of many of the individual organs of the digestive system, including crop, small intestine, and large intestine in both poults and chicks. Knight and Dibner (1998) also observed reduced intestine, liver, and pancreas growth following a delay in feeding of 3 days. Reduced residual yolk sac utilization has been reported to be associated with delayed access to feed (Noy and Sklan, 1998, 1999; Sklan and Noy, 2000). Morphological examination of the GI tract from birds held with no access to feed or water for 48 hours showed decreased villus growth; further, the major reduction in villus size and enterocytes per villus was not apparent until some 5 to 6 days later (Noy et al, 2001). Work has been published that indicates that the reduction in intestinal weight is associated with reduced enterocyte proliferation, increased enterocyte apoptosis, and reduced villus length and surface area (Noy et al., 2001; Noy and Sklan, 1998; Potturi et al., 2005; Uni et al., 1998). In addition, negative effects of fasting on intestinal protein deposition, brush border enzyme development, pancreatic enzyme secretion, and nutrient transport (Sklan and Noy, 2000, 2003; Uni et al., 2003b) have been reported. The mechanism responsible for these changes has not been determined but Geyra et al., (2002) provided evidence that a 48 hour fast reduced the expression of the cdxA and cdxB transcription factors, which could lead to negative effects on enterocyte maturation. Uni et al. (2003) reported changes in mucin secretion associated with delayed access to feed and Smirnov et al. (2006) reported increased mucin gene expression and intestinal mucin content following in ovo feeding of carbohydrates. In addition, other indicators of metabolic status, such as the shift to glycolytic metabolism in the liver, and the absorption of hydrophilic nutrients such as glucose and methionine, have been shown to be delayed by fasting (Noy and Sklan, 1999; Sklan and Noy, 2003b; Turner et al., 1999). Batal and Parsons (2002) reported a significantly higher ME at day 21 in birds fed Oasis for 24 or 48 hr at the time of hatch compared to fasted controls. This represents a long term change in nutrient utilization associated with a brief delay in nutrient availability. All of these factors could lead to a depression in digestive system function due to a reduction in absorption capability, and this may be an explanation for the long-term reduced growth described. Protection of the animal against invasion by opportunistic pathogens from the GI luminal microflora is due to both innate and adaptive immune systems. The GI tract epithelium plays a key role in maintaining this barrier because of the tight junctions that join one cell to another and the layer of mucin secreted by the goblet cells (Sklan, 2005). In addition, the epithelium itself has receptors that recognize pathogen-associated molecular patterns (Ishikawa et al., 2005). This non-specific binding provokes the synthesis of IL-8 (Berin et al., 1999) by the epithelial cells, leading to the initiation of an innate immune response involving heterophils and macrophages. If this layer of defense is not successful in preventing invasion, many intraepithelial leukocytes, including macrophages, dendritic cells and natural killer cells, populate the intestinal epithelium and respond to any foreign organism or antigen (Smirnov et al., 2004). Although the innate immune system is functional at hatch, it continues to develop over the first week of life (Bar-Shira and Friedman, 2006). The Bursa of Fabricius, the primary immune organ for Blymphocyte proliferation and differentiation in the vertebrate class Aves (Leslie, 1975; Peault et al., 1987), is found on the dorsal surface of the large intestine and is connected to the lumen of the intestine by the presence of the bursal duct (Schaffner et al., 1974). The GI tract itself is also heavily populated with both B- and T-lymphocytes. These can be found in the cecal tonsils, Meckels Diverticulum and scattered in the epithelium lining the intestinal lumen (Jeurissen et al., 1989; Lillehoj and Chung, 1992). The adaptive immune system, consisting of the products, cellular and molecular, of the clonal proliferation of lymphocytes in response to specific antigens, is largely undeveloped at hatch and is greatly affected by the nutrients and antigens it encounters during the first week of life (Bar-Shira et al., 2003).

56

Several researchers have reported effects of delayed feeding on the development of the immune system and disease resistance. Early work described effects on immune organ weights and lymphocyte proliferation (Dibner et al., 1998b). This was found to be associated with delayed IgA production, germinal center formation, and reduced resistance to a coccidiosis challenge (Dibner et al., 1998b; Dibner and Knight, 2003). The effects have been further explored in more recent work showing that the humoral immune response, particularly in the hindgut, and cellular immunological variables were lower in the feeddeprived birds (Bar-Shira et al., 2005; Juul-Madsen et al., 2004). Juul-Madsen et al. (2004) demonstrated reductions in relative expression of Major Histocompatiblity Complex class II antigens and of BU-1 cell surface antigen on B-lymphocytes. Bar-Shira et al. (2005) found reductions in CD4+ CD8+ double positive cells, reduced numbers of BU-1 positive cells and reduced expression of chIL-2 and CD3 mRNA, particularly in the hindgut. When birds were exposed to an antigen via the hindgut route, fasted birds showed a significantly depressed antibody response (Bar-Shira et al., 2005). These findings may help to explain the reduced disease resistance seen in neonatal birds denied access to feed. Effects of delayed access to feed are not limited to the GI and immune systems. There are also many indications that muscle itself is affected by early posthatch starvation. Indirect effects may be associated with the fact that the GI and immune tissues use glucose and glutamine as preferred energy substrates (Watford et al., 1979). Since the residual yolk has no remaining glycogen, if feed is not provided the bird must degrade protein, particularly skeletal muscle, for the amino acids used as substrates for gluconeogenesis. Glucogenic amino acids include glutamine, arginine and even limiting amino acids such as methionine. Thus, the nutrient profile of feed for the immediate post hatch period should be rich in highly available carbohydrates. In addition, there are direct effects of delayed feeding on skeletal muscle accretion. Effects of delayed feeding on muscle include reduced satellite cell proliferation (Mozdziak et al., 2002b), increased satellite cell apoptosis (Pophal et al., 2003), ultimately resulting in a reduction in breast yield at the time of marketing (Berri et al., 2006; Halevy et al., 2003). An increase in apoptosis in the GI system has also been reported (Potturi et al., 2005) and other support and demand organ systems should be examined to understand the degree to which apoptosis plays a role in organ weight changes associated with delayed access to feed. Another developmental process that could be altered by not feeding the hatchling is the timely delivery of microorganisms to the gut along with the substrates they need to survive. Delaying this process affects both their ability to become members of the resident microbial community, and also the ability of the bird to prevent them from invading the rest of the body. Under normal circumstances, hatchling birds appear to have a sterile GI lumen until it becomes populated by bacteria from hatchery waste or other hatchery-related microbial populations (Apajalahti, 2005; Pedroso et al., 2005). Provision of nutrients, particularly carbohydrates, is necessary to encourage colonization by saccharolytic organisms rather than protein fermenting putrefaction organisms (Apajalahti 2005). The microflora introduced with the feed or provided within the feed as a probiotic have several effects on GI development, including stimulation of innate immunity (Farnell et al., 2006; Haghighi et al., 2006), development of tolerance or response by the adaptive immune system (Klipper et al., 2001), stimulation of intestinal barrier function through mucin synthesis (Smirnov et al., 2004), modulation of epithelial differentiation (Furuse et al., 1991; Muller et al., 2005) and other effects related to the extraction of nutrients from otherwise unavailable ingredients such as cellulose (Apajalahti, 2005; Tellez et al., 2006). The influence of the intestinal microbiota on the host is currently a very active area of research that will no doubt reveal other aspects of this ancient relationship between bacteria and vertebrates. In summary, early feeding research has revealed that delayed access to feed can have developmental effects on the GI system that can result in secondary effects on performance. The balance of nutrients has not been extensively studied, although several researchers have observed that carbohydrates are crucial for both GI structural and functional development. The increased digestibility of saturated fats is thought to be associated with immature bile salt synthesis and entero-hepatic recycling. Finally, immediate access to feed has been shown to be essential for development of immune tolerance to nutrients and immune 57

response to pathogens. The role of enteral nutrition in the development and maintenance of the barrier function of the intestinal epithelium suggests that feed intake may be key for establishment of a stable microflora. References Apajalahti, J. 2005. Comparative gut microflora, metabolic changes, and potential opportunities. J. Appl. Poult. Res. 14:444-453. Bar-Shira, E., and A. Friedman. 2005. Ontogeny of gut associated immune competence in the chick. Israel J.Vet. Med. 60:42-50. Bar-Shira, E., and A. Friedman. 2006. Development and adaptations of innate immunity in the gastrointestinal tract of the newly hatched chick. Devel. Comp. Immun. 30:930-941. Bar-Shira, E., D. Sklan and A. Friedman. 2003. Establishment of immune competence in the avian GALT during the immediate post-hatch period. Devel. Comp. Immun. 27:147-157. Bar-Shira, E., D. Sklan and A. Friedman. 2005. Impaired immune responses in broiler hatchling hindgut following delayed access to feed. Vet. Immun. Immunopath. 105:33-45. Baranyiova, E. 1972. Influence of deutectomy, food intake, and fasting on the digestive tract dimensions in chickens after hatching. Acta Vet. Brno 41:373-384. Baranyiova, E., and J. Holman. 1976. Morphological changes in the intestinal wall in fed and fasted chickens in the first week after hatching. Acta Vet. Brno 45:151-158. Batal, A., and C. Parsons. 2002. Effects of age on nutrient digestibility in chicks fed different diets. Poult. Sci. 81:400-407. Batal, A., and C. Parsons. 2004. Utilization of various carbohydrate sources as affected by age in the chick. Poult. Sci. 83:1140-1147. Berin, M., D. McKay and M. Perdue. 1999. Immune-epithelial interactions in host defense. Am. J. Trop. Med. Hyg. 60:16-25. Berri, C., E. Godet, N. Hattab and M. Duclos. 2006. Growth and differentiation of the chicken Pectoralis major muscle: Effect of genotype and early nutrition. Arch. Tierz, Dummerstorf 49:31-32. Buddington, R. 1992. Intestinal nutrient transport during ontogeny of vertebrates. Am. J. Physiol. 32:R503-509. Buddington, R., and J. Diamond. 1989. Ontogenetic development of intestinal nutrient transporters. Ann. Rev. Physiol. 51:601-619. Dibner, J. 2005. Early nutrition of zinc and copper in chicks and poults: Impact on growth and immune function. Pages 23-32. In: Proc. Mid-Atlantic Nutrition Conference, Timonium, MD. Dibner, J., and C. Knight. 2001. Early feeding and nutritional programming in hatchling poultry. Pages 19. In: Proc. Arkansas Nutrition Conference, Fayetteville, AR. Dibner, J., and C. Knight. 2003. Early nutrition and immune development. Pages 172-178. In: Proc. California Animal Nutrition Conference, Fresno, CA. Dibner, J., C. Knight and F. Ivey. 1998a. The feeding of neonatal poultry. World Poultry 14:36-40. Dibner, J., C. Knight, M. Kitchell, C. Atwell, A. Downs and F. Ivey. 1998b. Early feeding and development of the immune system in neonatal poultry. J. Appl. Poult. Res. 7:425-436. Farnell, M., A. Donoghue, F. de los Santos, P. Blore, B. Hargis, G. Tellez and D. Donoghue. 2006. Upregulation of oxidative burst and degranulation in chicken heterophils stimulated with probiotic bacteria. Poult. Sci. 85:1900-1906. Ferket, P. 2006. Incubation and in ovo nutrition affects neonatal development. Pages 18-28. In: Annual Carolina Poultry Nutrition Conference Proc., Research Triangle Park, NC. Foye, O., Z. Uni and P. Ferket. 2006. Effect of feeding egg white protein, hydroxy-methylbutyrate and carbohydrates on glycogen status and neonatal growth in turkeys. Poult. Sci. 85:1185-1192. Freeman, B., and R. Vince. 1974. Development of the Avian Embryo. Chapman and Hall, London. Furuse, M., S. Yang, N. Niwa and J. Okumura. 1991. Effect of short chain fatty acids on the performance and the intestinal weight in germ free and conventional chicks. Brit. Poult. Sci. 32:159-165.

58

Garcia, A., and A. Batal. 2005. Changes in digestible lysine and sulfur amino acids needs of broiler chicks during the first three weeks posthatching. Poult. Sci. 84:1350-1355. Geyra, A., Z. Uni, O. Gal-Garber, D. Guy and D. Sklan. 2002. Starving affects CDX gene expression during small intestinal development in the chick. J. Nutr. 132:911-917. Gomez, M., and D. Polin. 1976. The use of bile salts to improve absorption of tallow in chicks, one to three weeks of age. Poult. Sci. 55:2189-2195. Haghighi, H., J. Gong, C. Gyles, M. Hayes, H. Zhou, B. Sanei, J. Chambers and S. Sharif. 2006. Probiotics stimulate production of natural antibodies in chickens. Clin. Vac. Imm. 13:975-980. Halevy, O., Y. Nadel, M. Barak, I. Rozinboim and D. Sklan. 2003. Early posthatch feeding stimulates satellite cell proliferation and skeletal muscle growth in turkey poults. J. Nutr. 133:1376-1382. Iji, P., A. Saki and D. Tivey. 2001. Body and intestinal growth of broiler chicks on a commercial starter diet. I. Intestinal weight and mucosal development. Brit. Poult. Sci. 42:505-513. Ishikawa, H., Y. Kanamori, H. Hamada and H. Kiyono. 2005. Development and function of organized gut associated lymphoid tissue. Elsevier, Burlington, MA. Jackson, S. 2005. Influence of pre-feeding a semi-solid hydrated supplement, OASIS, on development and performance of turkey poults. Masters Thesis. North Carolina State University, Raleigh, NC. Jeurissen, S., E. Janse, G. Koch and G. De Boer. 1989. Postnatal development of mucosa-associated lymphoid tissues in chickens. Cell Tis. Res. 258:119-124. Juul-Madsen, H., G. Su and P. Sorensen. 2004. Influence of early or late start of first feeding on growth and immune phenotype of broilers. Brit. Poult. Sci. 45:210-222. Kidd, M.T., M.A. Qureshi, P.R. Ferket and L.N. Thomas. 2000. Turkey Hen Zinc Source Affects Progeny Immunity and Disease Resistance. J. Appl. Poult. Sci. 9:414-423. Klipper, E., D. Sklan and A. Friedman. 2001. Response, tolerance or ignorance following oral exposure to a single dietary protein antigen in Gallus domesticus. Vaccine 19:2890-2897. Knight, C., and J. Dibner. 1998. Nutritional programming in hatchling poultry: Why a good start is important. Poultry Digest 57:20-26. Krogdahl, A., and J. Sell. 1989. Influence of age on lipase, amylase and protease activities in pancreatic tissue and intestinal contents of young turkeys. Poult. Sci. 68:1561-1568. Leslie, G. 1975. Ontogeny of the chicken humoral immune system. Am. J. Vet. Res. 36:482-485. Lilburn, M. 1998. Practical aspects of early nutrition for poultry. J. Appl. Poult. Res. 7:420-424. Lillehoj, H., and K. Chung. 1992. Postnatal development of T-lymphocyte subpopulations in the intestinal intraepithelium and lamina propria in chickens. Vet. Immunol. Immunopath. 31:347-360. McDowell, L. 2000. Vitamin D. Pages 200-252. In: Vitamins in Animal and Human Nutrition. Iowa State University Press, Ames, IA. Mikec, M., Z. Bidin, A. Valentic, V. Savic, T. Zelenika, R. Raguz-Duric, I. Novak and M. Balenovic. 2006. Influence of environmental and nutritional stressors on yolk sac utilization, development of chicken gastrointestinal system and its immune status. World's Poult. Sci. 62:31-40. Mozdziak, P., J. Evans and D. McCoy. 2002a. Early posthatch starvation induces myonuclear apoptosis in chickens. J. Nutr. 132:901-903. Mozdziak, P., T. Walsh and D. McCoy. 2002b. The effect of early posthatch nutrition on satellite cell mitotic activity. Poult. Sci. 81:1703-1708. Muller, C., L. Autenrieth and A. Peschel. 2005. Innate defenses of the intestinal epithelial barrier. Cell. Mol. Life Sci. 62:1297-1307. Nitsan, Z., G. Ben-Avraham, Z. Zoref and I. Nir. 1991. Growth and development of the digestive organs and some enzymes in broiler chicks after hatching. Br. Poult. Sci. 32:515-523. Noy, Y., A. Geyra and D. Sklan. 2001. The effect of early feeding on growth and small intestine development in the posthatch poult. Poult. Sci. 80:912-919. Noy, Y., and D. Sklan. 1995. Digestion and absorption in the young chick. Poult. Sci. 74:366-373. Noy, Y., and D. Sklan. 1998. Metabolic responses to early nutrition. J. Appl. Poult. Res. 7:437-451. Noy, Y., and D. Sklan. 1999. Energy utilization in newly hatched chicks. Poult. Sci. 78:1750-1756.

59

Noy, Y., and D. Sklan. 2002. Nutrient use in chicks during the first week posthatch. Poult. Sci. 81:391399. Noy, Y., and D. Sklan. 2004. Effects of metabolizable energy and amino acid levels on turkey performance from hatch to marketing. J. Appl. Poult. Res. 13:241-252. Noy, Y., Z. Uni and D. Sklan. 1996. Routes of yolk utilization in the newly hatched chick. Poul. Sci. 75S:13. Ohlsson, T., and H. Smith. 2001. Early nutrition causes persistent effects on pheasant morphology. Physiol. Biochem. Zool. 74:212-218. Peault, B., F. Dieterlen-Lievre and N. Le Dourarin. 1987. Cellular interactions occurring during primary lymphoid ontogeny in birds. Pages 39-64. In: Avian Immunology: Basis and Practice. Toivanen, A. and P. Toivanen (eds.). CRC Press, Boca Raton, FL. Pedroso, A., J. Menten and M. Lambais. 2005. The structure of bacterial community in the intestines of newly hatched chicks. J. Appl. Poult. Res. 14:232-237. Peebles, E., J. Croom, W. Maslin, S. Witmarsh, L. Daniel and I. Taylor. 2001. In ovo peptide YY and epidermal growth factor administration and their effects on growth and yolk utilization in neonatal meat-type chickens (Gallus domesticus). Comp. Biochem. Physiol. Part A 130:741-749. Pophal, S., J. Evans and P. Mozdziak. 2003. Myonuclear apoptosis occurs during early posthatch starvation. Comp. Biochem. Physiol. Part B 135:677-681. Potturi, P., J. Patterson and T. Applegate. 2005. Effects of delayed placement on intestinal characteristics in turkey poults. Poult. Sci. 84:816-824. Romanoff, A. 1960. The digestive system. Pages 429-532. In: Avian Embryo. Macmillan Company, New York, NY. Schaffner, T., J. Mueller, M. Hess, H. Cottier, B. Sordat and C. Ropke. 1974. The bursa of Fabricius: A central organ providing for contact between the lymphoid system and intestinal content. Cell Immun. 13:304-312. Sell, J., C. Angel, J. Piquer, E. Mallarino and H. Al-Batshan. 1991. Developmental patterns of selected characteristics of the gastrointestinal tract of young turkeys. Poult. Sci. 70:1200-1205. Sklan, D. 2001. Development of the digestive tract of poultry. World's Poult. Sci. 57:415-428. Sklan, D. 2003. Fat and carbohydrate use in posthatch chicks. Poult. Sci. 82:117-122. Sklan, D. 2005. Development of defense mechanisms in the digestive tract of the chick. J. Appl. Poult. Res. 14:437-443. Sklan, D., A. Geyra, E. Tako, O. Gal-Gaber and Z. Uni. 2003. Ontogeny of brush border carbohydrate digestion and uptake in the chick. Brit. J. Nutr. 89:747-753. Sklan, D., and Y. Noy. 2000. Hydrolysis and absorption in the small intestines of posthatch chicks. Poult. Sci. 79:1306-1310. Sklan, D., and Y. Noy. 2003a. Crude protein and essential amino acid requirements in chicks during the first week post hatch. Brit. Poult. Sci. 44:266-274. Sklan, D., and Y. Noy. 2003b. Functional development and intestinal absorption in the young poult. Brit. Poult. Sci. 44:651-658. Smirnov, A., D. Sklan and Z. Uni. 2004. Mucin dynamics in the chick small intestine are altered by starvation. J. Nutr. 134:736-742. Smirnov, A., E. Tako, P. Ferket and Z. Uni. 2006. Mucin gene expression and mucin content in the chicken intestinal goblet cells are affected by in ovo feeding of carbohydrates. Poult. Sci. 85:669-673. Sulaiman, A., E. Peebles, T. Pansky, T. Kellogg, W. Maslin and R. Keirs. 1996. Histological evidence for a role of the yolk stalk in gut absorption of yolk in the post-hatch broiler chick. Poult. Sci. 75S:48. Tako, E., P. Ferket and Z. Uni. 2004. Effects of in ovo feeding of carbohydrates and beta-hydroxy-betamethylbutyrate on the development of the chicken intestine. Poult. Sci. 83:2023-2028. Tako, E., P. Ferket and Z. Uni. 2005. Changes in chicken intestinal zinc transporter mRNA expression and small intestine functionality following intra-amniotic zinc-methionine administration. J. Nutr. Biochem. 15:339-346.

60

Tellez, G., S. Higgins, A. Donoghue and B. Hargis. 2006. Digestive physiology and the role of microorganisms. J. Appl. Poult. Res. 15:136-144. Turner, K., T. Applegate and M. Lilburn. 1999. Effect of feeding high carbohydrate or fat diets. 2. Apparent digestibility and apparent metabolizable energy of the posthatch poult. Poult. Sci. 78:15811587. Uni, Z., and P. Ferket. 2004. Methods for early nutrition and their potential. World's Poult. Sci. 60:101111. Uni, Z., S. Ganot and D. Sklan. 1998. Posthatch development of mucosal function in the broiler small intestine. Poult. Sci. 77:75-82. Uni, Z., Y. Noy and D. Sklan. 1999. Posthatch development of small intestinal function in the poult. Poult. Sci. 78:215-222. Uni, Z., A. Smirnov and D. Sklan. 2003. Pre- and posthatch development of goblet cells in the broiler small intestine: Effect of delayed access to feed. Poult. Sci. 82:320-327. Watford, M., P. Lund and H. Krebs. 1979. Isolation and metabolic characteristics of rat and chicken enterocytes. Biochem. J. 178:589-596. Yi, G., G. Allee, C. Knight and J. Dibner. 2005. Impact of glutamine and Oasis hatchling supplement on growth performance, small intestinal morphology and immune response of broilers vaccinated and challenged with Eimeria maxima. Poult. Sci. 84:283-293.

61

FEEDING THE HEN FOR OFFSPRING PRODUCTIVITY M. T. Kidd Mississippi State University Department of Poultry Science Box 9665 Mississippi State, MS 39762-9665 Phone: 662-325-3416 Fax: 662-325-8292 Email: mkidd@poultry.msstate.edu Summary Research presented herein on fat soluble vitamins and trace elements points to their importance in terms of dietary levels and sources on offspring performance. The fat soluble vitamins D and E improve offspring health, bone density and quality, and subsequent broiler growth. Maternal nutrition in this area should be investigated further as benefits may arise based on the chicks preferential absorption of glucose rather than fatty acids which may limit chick fat soluble vitamin absorption post hatch. The trace minerals zinc, manganese, and selenium have numerous metabolic needs and have been shown to improve chick quality as mediated via improved chick health. Zinc is involved in numerous enzyme and hormonal functions and is critical for normal cellular development, and subsequent tissue (cartilage, bone, and muscle) growth. Trace mineral interactions and specific levels and sources that affect offspring are still unclear. These proceedings outline micronutrient nutrition and make reference to improved chick quality in offspring. Introduction Poultry breeding stock must be fed adequate levels of micronutrients for proper yolk assimilation of vitamins and minerals. This is more critical in poultry, as opposed to animals that have embryos that develop in the womb, because nutritional adequacies at a consistent level are critical for optimal offspring. If one reviews the literature, numerous papers dealing with offspring abnormalities via maternal micronutrient deficiencies are present. Of concern to breeder nutritionists in integrations, however, is the ability to heighten hen micronutrient nutrition to optimize chick quality and subsequent broiler performance and yield. Discussions presented herein will focus on how maternal nutrition impacts metabolism and health of offspring to elicit the former effects. Attention for discussion will be centered around trace metals (zinc, manganese, and selenium) and the fat soluble vitamins (A, D, and E), and Lcarnitine. Trace Mineral Premix Considerations Key papers on trace mineral nutrition in hen diets that impact offspring performance are presented in Table 1.

This is Journal Article Number PS-11082 from the Mississippi Agricultural and Forestry Experiment Station supported by MIS-322230. Use of trade names in these proceedings publication does not imply endorsement by the Mississippi Agricultural and Forestry Experiment Station of the products, nor similar ones not mentioned.

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

62

Table 1. Trace mineral maternal effects on offspring. Bird type1 LB

Mineral Zinc

Supplemented level2 0, 10, 20, and 40 mg/kg Zn 0, 40, or 80 mg/kg of zinc methionine added to the control diet (100 mg/kg Zn) 0, 20, 200, and 2,000 mg/kg Zn 0 and 72 mg/kg Zn (oxide and methionine sources) 0 and 40 mg/kg Zn (oxide and methionine sources) 160 mg/kg Zn (sulfate and Availa Zn forms) 160 mg/kg Zn (sulfate and Availa Zn forms) 0, 75 mg/kg Zn (sulfate and amino acid form), and 80 mg/kg Mn (sulfate and amino acid form)

Progeny response3 The corn/soy based diet (28 ppm Zn and 0 ppm supplemented Zn) was adequate to support chick growth up to 3 wk posthatching, but caused some feather fraying. Chicks from hens fed added zinc methionine had improved survival to an E. coli challenge.

Reference Stahl et al. (1986)

Zinc

LB

Flinchum et al. (1989)

Zinc Zinc

LB BB

Mineral (Zn, Cu, and Fe) levels of tibia, liver, and pancreas and growth 3 wk posthatching were not affected by treatments. Progeny from Zn-Met treated parents had heightened cellular immunity and dry bone weight over ZnO and controls. No differences were observed for chick growth, bone ash and Zn, and humoral immunity. Supplemental Zn-O and Zn-Met increased humoral and cellular immunity, respectively, over control chicks, but differences in hatching chick weight did not occur. Progeny chick weight, organ weights, yolk sac weight, intestinal weights, and glucose metabolism were not affected by hen zinc. Progeny live performance was not impacted by hen zinc. Humoral immunity was increased in progeny from hens fed sulfate zinc. Chicks from parents fed Zn- and Mn-amino acid complexes had improved livability at Days 0 to 17 and 0 to 34. Growth and carcass responses were unaffected by breeder treatment.

Stahl et al. (1990) Kidd et al. (1992)

Zinc

BB

Kidd et al. (1993)

Zinc Zinc Zinc and manganese

BB BB BB

Hudson et al. (2004a) Hudson et al. (2004b, 2005) Virden et al. (2003)

63

Table 1 continued Bird type1 BB

Mineral Zinc and manganese

Supplemented level2 0, 75 mg/kg Zn (sulfate and amino acid form), and 80 mg/kg Mn (sulfate and amino acid form) 0, 75 mg/kg Zn (sulfate and amino acid form), and 80 mg/kg Mn (sulfate and amino acid form) 0 and 2 mg/kg Se 0 and 0.03, and 0 and 0.1 mg/kg Se

Progeny response3 Supplemental Zn and Mn from amino acid complexes improved cellular immunity and all treatments containing supplemental Zn and Mn resulted in increased thymus weights. Differences in bursal weights were inconsistent and humoral immunity differences did not occur. Progeny from hens fed supplemental Zn- and Mn-amino acid complexes had higher left and total ventricular mass than controls. Differences in ascites incidence, hematocrit values, and pulse oximetry did not occur. Se improved the growth rate of chicks. Supplementing the hen diets with 0.03 ppm Se increased growth 14 d posthatching. The addition of 0.1 ppm Se reduced, but did not prevent exudative diathesis in progeny fed the vitamin E adequate diet.

Reference Virden et al. (2004)

Zinc and manganese

BB

Virden et al. (2004)

Selenium Selenium

LB LB

Poley et al. (1941) Cantor and Scott (1974)

BB=broiler breeder hen. LB=leghorn breeder hen. Levels refer to diets unless otherwise noted. Test diet basal levels of minerals were as follows: Stahl et al. (1986), Flinchum et al., 1989, 100 ppm Zn; 28 ppm Zn; Stahl et al. (1990), 28 ppm Zn; Kidd et al. (1992), 72 ppm Zn; Kidd et al. (1993), 72 and 83 ppm Zn; Virden et al. (2003; 2004), 75 and 83 ppm of Zn and Mn, respectively; Cantor and Scott (1974), 0.02 ppm Se. 3 Progeny results represent treated parents. Progeny were fed common diets unless otherwise noted.
2

64

Zinc containing metalloenzymes are represented in all six enzyme classes. In addition to enzymes, zinc functions are integral for RNA and DNA synthesis and metallothionines. Further, hormone functions associated with muscle and bone deposition are dependent upon zinc. Insulin-like growth factor-1 has been showed to have reduced serum levels in animals fed zinc deficient diets. As the hen must incorporate all nutrients in ovo prior to lay, it is critical that circulating levels of minerals (i.e., zinc) are adequate for proper incorporation into egg for subsequent embryo and chicks development. Zinc levels fed to leghorn breeders resulted in no effect on growth of progeny (Stahl et al., 1986, 1990), whereas organic zinc and zinc and manganese supplementation to zinc and manganese adequate diets resulted in improved immunity, but not growth (Kidd et al., 1992, 1993; Virden et al., 2004). Research has indicated that the improvements in chick immunity as a result of mineral fortification of hen diets may result in improved livability. For example, Flinchum et al. (1989) demonstrated that leghorn breeders fed supplemental zinc methionine to a zinc adequate diet had progeny with improved survival to an E. coli challenge. Hudson et al. fed Cobb 500 breeder hens a diet containing zinc premix level of 160 ppm from either zinc sulfate or Availa zinc-amino acid complex from placement to termination (65 weeks of age) of the hen flock (Hudson et al., 2004a,b; 2005). Chick weight and subsequent chick performance were not affected by hen zinc. An experiment assessing metabolic needs of progeny post hatch revealed that progeny chick weight, organ weights, yolk sac weight, intestinal weights, and glucose metabolism were not affected by hen zinc (Hudson et al., 2004a). In one case, progeny from hens fed zinc sulfate had higher primary antibody response to a T dependent antigen at 15 days of age (Hudson et al., 2004b). Recently, Virden et al. (2003) demonstrated that Cobb 500 breeders fed supplemental zinc and manganese amino acid complexes had progeny with improved livability. Interestingly, an evaluation of the yield data of the former trial (Virden et al., 2003) indicated that hens fed zinc and manganese amino acid complexes had progeny with higher breast meat yield than that of hens fed lowers levels of inorganic zinc and manganese sources, pointing to the importance of these metal impacting broiler protein deposition when fed maternally. In addition to level, hen nutritionists should consider mineral source availability (i.e., zinc and manganese) in order to heighten progeny immunity and subsequent chick health. Hence, organic minerals are more available than inorganic forms deeming potential incorporation into yolk as greater than sulfates or oxides. Selenium is typically noted as the only mineral that impacts chick growth as a result of parental supplementation (Poley et al., 1941; Cantor and Scott, 1974) and its effect has been measured up to day 14 in chicks (Cantor and Scott, 1974). The impact of selenium on antioxidant status is referenced with vitamin E in the next section. Soluble Vitamin and Vitamin-Like Compound Premix Considerations Key references are listed in Table 2 with reference to research in hens dealing with fat soluble vitamins and L-carnitine effects on offspring performance. Vitamin E has large impact on progeny with respect to increasing levels in yolk and subsequent effects on chick health and antioxidant status. Commercial broiler breeders fed supplemental vitamin E levels had chicks with improved lymphocyte proliferation (Haq et al., 1996), improved humoral immunity (Haq et al., 1996; Boa-Amponsem et al., 2001), and reduced susceptibility to peroxidation (Surai et al., 1999; Surai, 2000). In addition, the combination of selenium and vitamin E to broiler breeders has been shown to increase liver glutathione activity of progeny (Surai, 2000). Also, the combination of vitamin E and -carotene improved lymphocyte proliferation (Haq et al., 1996). However, research has indicated that carotenoids do not positively impact chick growth or immunity when fed to hens (Haq et al., 1995) and are not absorbed (yolk to embryo to chick) well by the chick (Haq and Bailey, 1996). Care must be taken when changing vitamin levels in the premix as vitamin A has been shown to decrease vitamin E availability to chicks (Combs, 1976; Surai et al., 1998).

65

Table 2. Fat soluble vitamin and L-carnitine effects on offspring. Bird type1 LB LB

Item Vitamin A Vitamin A

Supplemented level 1.5 X 106 vitamin A/kg of diet (retinyl palmitate) 0, 3, 30, and 120 g/g diet of vitamin A 0.02% -carotene, canthaxanthin, or lutein 0.02% -carotene, canthaxanthin, or lutein -carotene, canthaxanthin, lutein, and -tocopherol acetate 0, 90, 150, 300, 450, and 900 mg of vitamin E/kg of diet 147 versus 365 g vitamin E/g of feed 10 and 300 IU/kg diet of - tocopherol acetate 0.2 or 0.4 mg/kg diet of Se and 40, 100, or 200 mg/kg diet of vitamin E

Progeny response2 At 1 d posthatching chicks fed high vitamin A had depleted plasma tocopherols, but normal glutathione activities Vitamin A in the hens diet increased vitamin A in embryonic liver and the chick, but decreased vitamin E, carotenoids, and ascorbic acid in embryonic and chick liver. The carotenoids fed to hens did not positively impact growth, immune organ weights, or humoral immunity in chicks 5 weeks posthatching. Results indicate that although carotenoids are transferred from the hen to the yolk, they are not absorbed well by the embryo and subsequent chick. Control and -carotene treatments resulted in chicks with improved feed:gain 3 weeks posthatching over other treatments. -carotene, vitamin E, and their combination improved lymphocyte proliferation, but only vitamin E improved humoral immunity. Vitamin E levels of 150 and 450 mg/kg, but not 90, 300, and 900 mg/kg, increased passively transferred antibody levels in chicks to Brucella abortus up to Day 7 of age. Increased vitamin E in the hens diet resulted in increased vitamin E levels in chicks yolk sac membrane, liver, brain, and lung. These tissues, especially brain, had reduced susceptibility to peroxidation. Increasing vitamin E in some breeder lines increased progeny antibody titers to sheep red blood cells at hatch. The combination of increased Se and vitamin E increased liver glutathione activity in chicks. Increasing Se increased Se dependent glutathione peroxidase in chick liver.

Reference Combs (1976) Surai et al. (1998)

Carotenoids Carotenoids

LB LB

Haq et al. (1995) Haq and Bailey (1996) Haq et al. (1996)

Carotenoids and vitamin E

BB

Vitamin E

LB

Jackson et al. (1978) Surai et al. (1999)

Vitamin E

BB

Vitamin E Vitamin E and selenium

BB BB

Boa-Amponsem et al. (2001) Surai (2000)

66

Table 2 continued Bird type1 LB

Item Vitamin D

Supplemented level 24, 2,500, and 5,000 g/kg diet of cholecalciferol D3 and 25-OHD3

Progeny response2 Differences in chick body weight did not occur at 2 or 4 weeks posthatching. Tibial Ca was increased at 2 weeks posthatching by the high dose of vitamin D3 and tibial ash was increased by both vitamin D3 levels at 4 weeks posthatching. Chicks fed 0.90% Ca and 40 g/kg of 25-OHD3 and from hens fed 4000 IU/kg of D3 had improved BW and tibia ash, and the lowest level of tibial dyschondroplasia and Ca rickets. Increasing 45 wk old hen D3 from 250 to 4,000 IU/kg of diet resulted in a linear increase body weight gain at d 16 in progeny. In some of the progeny growout experiments, hens fed L-carnitine had reduced carcass abdominal fat. An interaction with hen dietary Lcarnitine and progeny high nutrient density resulted in improved breast meat.

Reference Ameenuddin et al. (1986)

Vitamin D

BB

Atencio et al. (2005a) Atencio et al. (2005b) Kidd et al. (2005)

Vitamin D L-Carnitine

BB BB

4,000 IU/kg of D3 0 and 25 mg/kg Lcarnitine

1 2

BB=broiler breeder hen. LB=leghorn breeder hen. Progeny results represent treated parents and progeny were fed common diets unless otherwise noted.

67

As the primary function of vitamin D is to increase intestinal absorption of calcium and phosphorus, attention should be made to hen dietary vitamin D level and source as improvements in early growth and bone quality have been observed (Atencio et al., 2005a,b). This was initially shown by Griminger et al. (1966) who showed that higher maternal vitamin D levels increased progeny bone mineralization. Recent research at the University of Georgia has been instrumental in demonstrating the impact of hen vitamin D on progeny growth and bone abnormalities (Atencio et al., 2005a,b; Driver et al., 2006) and hen vitamin D metabolites on the former (Atencio et al. 2005c). The work at the University of Georgia has demonstrated that hens fed between 2,000 and 4,000 IU vitamin D3/kg of diet have progeny with optimal bone quality and tibia ash, but vitamin D3 levels above 4,000 IU may be need to optimize progeny growth. We recently fed broiler breeder hens (beginning at 21 weeks of age) diets with and without Lcarnitine (0 of 25 mg/kg of diet) (Kidd et al., 2005). Progeny performance and carcass traits were evaluated in three hatches (30, 35, and 37 weeks). Hen dietary carnitine reduced abdominal fat pad in progeny at processing-irrespective of progeny diet. Also, hen carnitine interacted with increased progeny dietary nutrient density resulting in increased breast meat accretion. Coupled with the former results concerning fat sources, it appears that hen ingredients and nutrients involved in fat metabolism can influence progeny carcass traits. Conclusions These proceedings which accompany a presentation at the 2007 Mid-Atlantic Nutrition Conference clearly demonstrate that the fat soluble vitamins (i.e., vitamins D and E) and the trace elements (i.e., zinc, manganese, and selenium) discussed herein impact offspring in a number of ways. Hence, the former nutrients impact offspring metabolism, health, and growth. Mention was made to another micronutrient, L-carnitine, which has been shown to impact offspring development. Poultry companies will continue to become more competitive in the future. As such, nutritional regimes, such as hen feeding to impact offspring, will be further evaluated to measure practicality and subsequent profitability. This is a difficult area to demonstrate cause and effect. For example, the specific zinc need of a hen is unknown, in addition to the fact that this need may vary depending on source of zinc and the parameter of interest. To further complicate the assessment of hen zinc needs is quantifying the need parameters of interest as various offspring criteria. There are numerous examples that make progeny assessment to quantify hen nutrition as complicated. Hence, is the need of higher levels of vitamin D and E for offspring a function of heightened requirements, or the fact the fatty acid absorption in the young chick, and thus fat soluble vitamin absorption, is suboptimal? Regardless of the difficulties in assessing the need of nutrients in hen diets for offspring performance, this area will continually be researched so as to optimize offspring health and chick quality. References
Ameenuddin, S., M.L. Sunde, H.F. DeLuca and M.E. Cook. 1986. Excessive cholecalciferol in a layers diet: Decline in some aspects of reproductive performance and increased bone mineralisation of progeny. British Poult. Sci. 27:671-677. Atencio, A., H. Edwards, Jr., and G. Pesti. 2005a. Effect of level of cholecalciferol supplementation of broiler breeder hen diets on the performance and bone abnormalities of the progeny fed diets containing various levels of calcium or 25-hydroxycholecalciferol. Poult. Sci. 84:1593-1603. Atencio, A., H. Edwards, Jr., and G. Pesti. 2005b. Effects of vitamin D3 dietary supplementation of broiler breeder hens on the performance and bone abnormalities of the progeny. Poult. Sci. 84:1058-1068. Atencio, A., G.M. Pesti and H.M. Edwards, Jr. 2005c. Twenty-five hydroxycholecalciferol as a cholecalciferol substitute in broiler breeder hen diets and its effect on the performance and general health of the progeny. Poult. Sci. 84:1277-1285. Boa-Amponsem, K., S.E. Price, P.A. Geraert, M. Picard and P.B. Siegel. 2001. Antibody responses of hens fed vitamin E and passively acquired antibodies of their chicks. Avian Diseases 45:122-127.

68

Cantor, A.H., and M.L. Scott. 1974. The effect of selenium in the hens diet on egg production, hatchability, performance of progeny and selenium concentration in eggs. Poult. Sci. 53:1870-1880. Combs, G.F., Jr. 1976. Differential effects of high dietary levels of vitamin A on the vitamin E-selenium nutrition of young and adult chickens. J. Nutrition 106:967-975. Driver, J.P., A. Atencio, G.M. Pesti, H.M. Edwards, Jr., and R.I. Bakalli. 2006. The effect of maternal dietary vitamin D3 supplementation on performance and tibial dyschondroplasia of broiler chicks. Poult. Sci. 85:39-47. Flinchum, J.D., C.F. Nockles and R.E. Moreng. 1989. Aged hens fed added zinc methionine had chicks with improved performance. Poult. Sci. 68(Suppl. 1):55 (Abstract). Griminger, P. 1966. Influence of maternal vitamin D intake on growth and bone ash of offspring. Poult. Sci. 45:849851. Haq, A., and C. Bailey. 1996. Time course evaluation of carotenoid and retinol concentrations in posthatch chick tissue. Poult. Sci. 75:1258-1260. Haq, A., C.A. Bailey and A. Chinnah. 1996. Effect of -carotene, canthaxanthin, lutein, and vitamin E on neonatal immunity of chicks when supplemented in the broiler breeder diets. Poult. Sci. 75:1092-1097. Haq, A., C.A. Bailey and A.D. Chinnah. 1995. Neonatal immune response and growth performance of chicks hatched from single comb white leghorn breeders fed diets supplemented with -carotene, canthaxanthin, or lutein. Poult. Sci. 74:844-851. Hudson, B.P., B.D. Fairchild, J.L. Wilson, W.A. Dozier III and R.J. Buhr. 2004a. Breeder age and zinc source in broiler breeder hen diets on progeny characteristics at hatching. J. Appl. Poult. Res. 13:55-64. Hudson, B.P., W.A. Dozier III, B.D. Fairchild, J.L. Wilson, J.E. Sander and T.L. Ward. 2004b. Live performance and immune responses of straight-run broilers: Influences of zinc source in broiler breeder hen and progeny diets and ambient temperature during the broiler production period. J. Appl. Poult. Res. 13:291-301. Hudson, B.P., W.A. Dozier III and J.L. Wilson. 2005. Broiler live performance response to dietary zinc source and the influence of zinc supplementation in broiler breeder diets. An. Feed Sci. Tech. 118:329-335. Jackson, D.W., G.R. Law and C.F. Nockels. 1978. Maternal vitamin E alters passively acquired immunity of chicks. Poult. Sci. 57:70-73. Kidd, M.T., C.D. McDaniel, E.D. Peebles, S.J. Barber, A. Corzo and S.L. Branton. 2005. Breeder hen dietary Lcarnitine affects progeny carcase traits. British Poult. Sci. 46:97-103. Kidd, M.T., N.B. Anthony and S.R. Lee. 1992. Progeny performance when dams and chicks are fed supplemental zinc. Poult. Sci. 71:1201-1206. Kidd, M.T., N.B. Anthony, L.A. Newberry and S.R. Lee. 1993. Effect of supplemental zinc in either a corn-soybean or a milo and corn-soybean meal diet on the performance of young broiler breeders and their progeny. Poult. Sci. 72:1492-1499. Poley, W.E., W.O. Wilson, A.L. Moxon and J.B. Taylor. 1941. The effect of selenized grains on the rate of growth in chicks. Poult. Sci. 20:171-179. Stahl, J.L., M.E. Cook and M.L. Sunde. 1986. Zinc Supplementation: Its effect on egg production, feed conversion, fertility and hatchability. Poult. Sci. 65:2104-2109. Stahl, J.L., J.L. Greger and M.E. Cook. 1990. Breeding-hen and progeny performance when hens are fed excessive dietary zinc. Poult. Sci. 69:259-263. Surai, P.F. 2000. Effect of selenium and vitamin E content of the maternal diet on the antioxidant system of the yolk and the developing chick. Brit. Poult. Sci. 41:225-243. Surai, P.F., I.A. Ionov, T.V. Kuklenko, I.A. Kostjuk, A. MacPherson, B.K. Speake, R.C. Noble and N.H.C. Sparks. 1998. Effect of supplementing the hens diet with vitamin A on the accumulation of vitamins A and E, ascorbic acid and carotenoids in the egg yolk and in the embryonic liver. Brit. Poult. Sci. 39:257-263. Surai, P.F., R.C. Noble and B.K. Speake. 1999. Relationship between vitamin E content and susceptibility to lipid peroxidation in tissues of the newly hatched chick. Brit. Poult. Sci. 40:406-410. Virden, W.S., J.B. Yeatman, S.J. Barber, K.O. Willeford, T.L. Ward, T.M. Fakler, R.F. Wideman, Jr., and M.T. Kidd. 2004. Immune system and cardiac functions of progeny from dams fed diets differing in zinc and manganese level and source. Poult. Sci. 83:344-351. Virden, W.S., J.B. Yeatman, S.J. Barber, C.D. Zumwalt, T.L. Ward, A.B. Johnson and M.T. Kidd. 2003. Hen mineral nutrition impacts progeny livability. J. Applied Poult. Res. 12:411-416.

69

THE INFLUENCE OF EARLY NUTRITION: WHEN IS DAY 1?


M.S. Lilburn Department of Animal Sciences Gerlaugh Hall Ohio State University/OARDC Wooster, OH 44691 Phone: 330-263-3992 FAX : 330-263-3949 Email: lilburn.1@osu.edu Summary Chicks and poults are dependent upon yolk lipids for a large proportion of their embryonic energy needs and at hatch are required to make an abrupt metabolic switch. The logistics of the poultry industry dictates that chicks are usually placed into growing facilities within 24 hours of hatch, but with poults, there is often a delay of 24 to 72 hours before the poults are housed with access to feed and water. It is assumed and continually stated that the residual yolk sac is a reserve source of nutrients for the hatchling during this early post-hatch period yet actual residual yolk composition data in the literature suggest otherwise. While delays in placement are a given, what is important to understand is the extent to which this results in a temporary or permanent reduction in overall growth and development. This hypothesis can only be tested if experiments include both chronological age comparisons and similar ages relative to access to feed and water. The objective of this paper is to review the literature on early nutrition, particularly fasting, and address it from the perspective that Day 1 is the first day with access to feed and water, not the day of hatch. It is commonly accepted that lipid digestion is poor in young chicks and poults and increases with age. When the actual digestibility of dietary lipids is determined in the first 7 days post-hatch in poults, however, it is observed that unsaturated fatty acids (18:2, 18:3) are highly digestible and this agrees with reports in the literature. This suggests that we may need to rethink our approach to what types and levels of fat may be best in our starting diets. The Status at Hatch The topic for this group of papers is early nutrition and the contributions that maternal carryover effects and early nutrition play in the overall course of growth and development. In the context of this discussion, it is important to first discuss the status of the newly hatched chick or poult and factors that may control the magnitude of the response to early nutritional intervention. It is well recognized that hen age is positively correlated with many aspects of reproductive performance including egg size and hatchability (Wilson, 1991) and the positive relationship between egg weight and chick or poult weight at hatch has also been well documented (Reinhart and Moran, 1979; Shanawany, 1987). The positive relationship between older hen age, egg size and hatchling weight is not simply due to a physically larger egg. Applegate and Lilburn (1996) took commercial turkey hatching eggs from the same flock at 5-6 week intervals from 36 to 55 weeks of age. There were the expected effects of hen age on egg size and hatchling weight but when the data from all ages was stratified to equal egg weight classes, hatchlings from the older hens still were consistently 3-4% heavier than those from younger hens. This supports the hypothesis that age associated maternal investment factors other than simply greater output of egg materials has a positive influence on overall embryonic development. Reidy et al. (1994) reported that over a 24-week production cycle, egg size in commercial turkey breeders increased 11% but the proportion of yolk increased 21% with only a 7% increase in albumen weight. In
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

70

the latter report by Applegate and Lilburn (1996), percentage yolk increased from 28 to 34% over the course of the production period whereas albumen weight declined from 60 to 53% during the same period. It is commonly inferred that at hatch, the residual yolk sac is an immediate source of nutrients to the hatchling as it makes the transition from yolk derived nutrients to commercial diets. This assumption is not supported by data in the literature, however, and the relative proportions of key yolk nutrients at hatch and their potential fate is well described by Dibner et al. (1998). Triglyceride accounts for about 50% of the lipid in the residual yolk sac (Donaldson, 1967) but this represents less than a gram of actual triglyceride at hatch in chicks (Noble and Moore, 1964) or poults (Ding et al., 1995). From an energetic standpoint, there is not sufficient residual neutral lipid to meet maintenance energy needs for more than a day (Akiba and Murakami, 1995; Lilburn, 1998). The residual yolk sac contains a considerable quantity of phospholipid and this may represent the key lipid pool within the residual yolk sac. The final steps of post-hatch resorption of the residual yolk sac are stimulated by feed consumption (Moran, 1989; Moran and Reinhart, 1980) and there can be active transport of yolk materials into the intestine up to 72 hrs posthatch (Esteban et al., 1991; Noy et al., 1996). Given the relative immaturity of the intestine at hatch (Dror et al., 1977; Sell et al., 1991), the phospholipid pool within the yolk sac may represent a physiologic reserve of residual fatty acids needed for the synthesis of new cellular membranes within proliferating enterocytes. This is supported by published observations that immediately post-hatch, proliferation within the intestine of chicks (Uni et al., 1998) and poults (Applegate et al., 1999a) occurs both along the villus and within the crypt region. In turkey poults, it has been reported that the carbohydrate status at hatch can be tenuous due to any number of factors that might accelerate the utilization of stored glycogen during the hatching process (Donaldson and Christensen, 1991; Donaldson et al., 1991; Christensen et al., 2001). This has been the rationale for the development of in ovo feeding strategies whose objective has been to improve the overall carbohydrate status of late term embryos and enhance post-hatch growth (Uni et al., 2005). This latter approach to post-hatch manipulation needs to be carefully monitored, however, because of potential interactions between age of the breeder flock and early post-hatch glucose regulatory mechanisms (Applegate and Lilburn, 1999; Applegate et al., 1999b). There are also reports in the literature showing that plasma glucose levels are tightly controlled even in chicks or poults that are exposed to 24 to 48 hr periods of post-hatch fasting. Turner et al. (1999a) reported that at 48 hours post-hatch, there were no differences in either liver glycogen or plasma glucose in poults that were fed immediately post-hatch or held without feed or water for the initial 48 hours. Houpt (1958) reported that in chicks, there were no changes in circulating glucose concentrations until at least two days into a prolonged fast. When is Day 1, the hatchery or the farm ? The simple question, When is day 1? is an important discussion point in the context of the considerable data being generated on early nutrition and the negative metabolic effects of variable periods of fasting on early growth of chicks and poults. Numerous authors have accurately described the variable period between hatch and placement on the farm as a rationale for studying the effects of post-hatch feed restriction on a host of early growth parameters (Dibner et al., 1998; Turner et al., 1999b; Mozdziak et al., 2002; Halevy et al., 2000, 2003; Moore et al., 2005; Velleman and Mozdziak, 2005). In almost all cases, however, the data collection was confined to days post-hatch rather than imposing a secondary treatment, day post-feeding. Turner et al. (1999a) studied the interactions between diets that were formulated to be high in available carbohydrate (CHO) or fat (FAT) and immediate or delayed access (DA) to feeding (48 hr). In these experiments, it was assumed that Day 1 was the first day with access to feed and water so all the data was presented as days post-feeding. In the initial two experiments, there were no differences in body weight between treatments at 12 days (Exp. 1) or 13 days (Exp. 2) post-feeding. The CHO diet improved early growth and feed intake through 5 days post-feeding but the FAT treatment improved both gain and

71

feed intake thereafter. In Exp. 2 and Exp. 3, liver glycogen and plasma glucose levels were significantly greater in the DA poults fed the CHO diets compared with the FAT diets at 2 days post-feeding and the plasma glucose levels stayed elevated through 5 days post-feeding. In Exp. 3, poults that had DA access to feed had elevated plasma glucose concentrations 30 and 60 minutes following a glucose challenge at both 4 and 7 days post-feeding. This suggests that metabolic homeostasis, at least with respect to carbohydrate metabolism had not yet been normalized. Dibner et al. (1998) compared several aspects of immune tissue development in fasted chicks compared with chicks fed a specialized hydrated nutritional supplement (OASISTM) for the first two days post-hatch. In this particular study, day of hatch was considered to be Day 0 and the subsequent days were considered Days 1 and 2. The weight of the bursa was consistently, though not always significantly, heavier in the chicks fed OASIS compared with the fasted controls through 10 days of age. From 10 to 15 and 15 to 20 days of age, however, there were considerable increases in bursal weight in both the OASIS and fasted chicks though the response was blunted in the fasted chicks. What is intriguing about this set of data is that the early treatment effects were most clearly expressed at least 10 days post-feeding (Figure 1) and this delayed response was even more dramatic with respect to the number of germinal centers in the cecal tonsil. This suggests that for some systems, the biological programming that occurs with respect to tissue function may be negatively influenced days or weeks after an insult, i.e. fasting, is removed. Figure 1. Bursa weight as a function of age in fasted birds or birds fed a hydrated nutritional supplement (HNS) for post-hatch Days 0, 1 and 2. Adapted from Dibner et al., 1998.

With the importance of breast meat to both the broiler and turkey industries, it is logical that early nutrition practices would be considered in light of maximizing the yield of these important muscles. Many of these studies have centered on the proliferation and differentiation of satellite cells, the pool of cells that through fusion with existing myofibers, are the source of post-hatch nuclear material during the process of muscle hypertrophy (Moss and Leblond, 1971; Campion, 1984.). Halevy et al. (2000; 2003) took a similar approach to Dibner et al. (1998) with chicks and poults in that half the birds were fed at hatch and the other treatment group was fasted for 48 hours prior to feeding. In the paper with chicks, the authors concluded that the initial fasting period resulted in a significant reduction in body weight through 41 days. The data in Figure 2 suggest that at 3 days post-

72

feeding (5 days of age for fasted chicks), there were no differences in body weight or relative breast muscle weight (g/100 g BW). In the paper with poults, the authors again tried to support a hypothesis that Figure 2. Body weight and breast muscle weight as a percentage of body weight of broiler chickens fed or starved during the first 2 d of life. Adapted from Halevy et al. (2000).

early fasting had a profound and prolonged effect on satellite cell proliferation but again, their data do not fully support that hypothesis. The data in Figure 3 show clearly that thymidine incorporation into satellite cells is delayed by early fasting but peak incorporation is almost the same in fed and fasted poults at similar time points post-feeding. With respect to total satellite cell numbers per gram of tissue, the peak in cell numbers is sharper and earlier in fed poults but the slightly lower but broader peak in the fasted poults again suggests that actual cell numbers may not be significantly decreased. This is similar to the data of Moore et al. (2005). In this study, poults were fasted for 72 hours and the body weight and satellite cell mitotic activity patterns were similar when one compares them at similar days post-feeding. Figure 3. Labeled thymidine incorporation into DNA in satellite cells (A)A and number of satellite cells per gram of breast muscle (B) of fed and 2 d feed-deprived (starved poults). Adapted from Halevy et al., 2003.

73

The data are shown in Figure 4. Velleman and Mozdziak (2005) presented data on proteoglycan expression in breast muscle from fasted chicks (72 hr). Their data also supports the hypothesis that onset of feeding, rather than simply fasting per se may play a bigger role in early muscle growth and development than previously thought. Figure 4. Body weight and satellite cell mitotic activity in pectoralis muscle in poults fed or fasted for 72 hours. Adapted from Moore et al., 2005.

In summary, studies addressing the importance of early nutrition should at least consider the effects of feeding as an important initiator of muscle growth and other systems as well. It is also important to remember that while fasting post-hatch is a reality in modern poultry production, 72 hours is an extreme and this further emphasizes the need for a biological control treatment, not simply chronological age comparisons. References Akiba, Y., and H. Murakami. 1995. Partitioning of energy and protein during early growth of broiler chicks and contribution of vitelline residue. Pages 45-52. In: Proc. 10th European Symposium on Poultry Nutrition. Applegate, T.J., and M.S. Lilburn. 1996. Independent effects of hen age and egg size on incubation and poult characteristics in commercial turkeys. Poult. Sci. 75:1210-1216. Applegate, T.J., and M.S. Lilburn. 1999. Effect of turkey (Meleagridis gallopavo) breeder hen age and egg size on poult development. 1. Intestinal growth and glucose tolerance of the turkey poult. Comp. Biochem. Physiol. Pt. B 124:371-380. Applegate, T.J., E. Ladwig, L. Weissert and M.S. Lilburn. 1999a. Effect of hen age on intestinal development and glucose tolerance of the Pekin duckling. Poult. Sci. 78:1485-1492. Applegate, T.J., J.J. Dibner, M.L. Kitchell, Z. Uni and M.S. Lilburn. 1999b. Effect of turkey (Meleagridis gallopavo) breeder hen age and egg size on poult development. 2. Intestinal villus growth, enterocytes migration and proliferation of the turkey poult. Comp. Biochem. Physiol. Pt. B 124:381-389. Campion, D.R. 1984. The muscle satellite cell: a review. Int. Rev. Cytol. 87:225-251. Christensen, V.L., M.J. Wineland, G.M. Fasenko and W.E. Donaldson. 2001. Egg storage effects on plasma glucose and supply and demand tissue glycogen concentrations of broiler embryos. Poult. Sci. 80:1729-1735.

74

Dibner, J., C.D. Knight, M.L. Kitchell, C.A Atwell, A.C. Downs and F.J. Ivey. 1998. Early feeding and development of the immune system in neonatal poultry. J. Appl. Poult. Res. 7:425-436. Ding, S.T., K.E. Nestor and M.S. Lilburn. 1995. The concentration of different lipid classes during late embryonic development in a randombred turkey population and a subline selected for increased body weight at sixteen weeks of age. Poult. Sci. 74:374-382. Donaldson, W.E. 1967. Lipid composition of chick embryo and yolk as affected by stage of incubation and maternal diet. Poult. Sci. 46:693-697. Donaldson, W.E., and V.L. Christensen. 1991. Dietary carbohydrate level and glucose metabolism in turkey poults. Comp. Biochem. Physiol. 98:347-350. Donaldson, W.E., V.L. Christensen and K.K. Krueger. 1991. Effects of stressors on blood glucose and hepatic glycogen concentrations in turkey poults. Comp. Biochem. Physiol. 100A:945-947. Dror, Y., I. Nir and Z. Nitsan. 1977. The relative growth of internal organs in light and heavy breeds. Br. Poult. Sci. 18:493-496. Esteban, S., M. Moreno, J.M. Rayo and J.A. Tur, 1991. Gastrointestinal emptying in the final days of incubation in the chick embryo. Br. Poult. Sci. 32:279-284. Halevy, O., A. Geyra, M. Barak, Z. Uni and D. Sklan. 2000. Early posthatch starvation decreases satellite cell proliferation and skeletal muscle growth in chicks. J. Nutr. 130:858-864. Halevy, O., Y. Nadel, M. Barak, I. Rozenboim and D. Sklan. 2003. Early posthatch feeding stimulates satellite cell proliferation and skeletal muscle growth in turkey poults. J. Nutr. 133:1376-1382. Houpt, T.R. 1958. Effects of fasting on blood sugar levels in baby chicks of varying ages. Poult. Sci. 37:1452-1459. Lilburn, M.S. 1998. Practical aspects of early nutrition for poultry. J. Appl. Poult. Res. 7:420-424. Moore, D.T., P.R. Ferket and P.E. Mozdziak. 2005. Early post-hatch fasting induces satellite cell selfrenewal. Comp. Biochem. Physiol. Pt. A. 142:331-339. Moran, E.T., Jr. 1989. Effects of posthatch glucose on poults fed and fasted during yolk sac depletion. Poult. Sci. 68:1141-1147. Moran, E.T., Jr., and B.S. Reinhart. 1980. Poult yolk sac amount and composition upon placement: effect of breeder age, egg weight, sex, and subsequent change with feeding or fasting. Poult. Sci. 59:15211528. Moss, F.P., and C.P. Leblond. 1971. Satellite cells as the source of nuclei in muscles of growing rats. Anat. Rec. 170:421-436. Mozdziak, P.E., T.J. Walsh and D.W. McCoy. 2002. The effect of early posthatch nutrition on satellite cell mitotic activity. Poult. Sci. 81:1703-1708. Noble, R.C., and J.H. Moore. 1964. Studies on the lipid metabolism of the chick embryo. Can. J. Biochem. 42:1729-1741. Noy, Y., Z. Uni and D. Sklan. 1996. Routes of yolk utilization in the newly-hatched chick. Br. Poult. Sci. 37:987-996. Reidy, R., J.L. Atkinson and S. Leeson. 1994. Strain comparisons of turkey egg components. Poult. Sci. 73:388-395. Reinhart, B.S., and E.T. Moran, Jr. 1979. Incubation characteristics of eggs from older small white turkeys with emphasis on the effects due to egg weight. Poult. Sci. 58:1599-1605. Sell, J.L., C.R. Angel, F.J. Piquer, E.G. Mallarino and H.A. Al-Batsham. 1991. Developmental patterns of selected characteristics of the gastrointestinal tract of young turkeys. Poult. Sci. 70:1200-1205. Shanawany, M.M. 1987. Hatching weight in relation to egg weight in domestic birds. W. Poult. Sci. J. 43:107-115. Turner, K.A., T.J. Applegate and M.S. Lilburn. 1999 a. Effects of feeding high carbohydrate or fat diets. 1. Growth and metabolic status of the posthatch poult following immediate or delayed access to feed. Poult. Sci. 78:1573-1580. Turner, K.A., T.J. Applegate and M.S. Lilburn. 1999 b. Effects of feeding high carbohydrate or fat diets. 2. Apparent digestibility and apparent metabolizable energy of the posthatch poult. Poult. Sci. 78:1581-1587.

75

Uni, Z., R. Platin and D. Sklan. 1998. Cell proliferation in chicken intestinal epithelium occurs both in the crypt and along the villus. J. Comp. Physiol. B 168:241-247. Uni, Z., P.R. Ferket, E. Tako and O. Kedar. 2005. In ovo feeding improves energy status of late-term chicken embryos. Poult. Sci. 84:764-770. Velleman, S.G. and P.E. Mozdziak, 2005. Effect of posthatch feed deprivation on heparin sulfate proteoglycan, syndecan-1, and glypican expression: Implications for muscle growth potential in chickens. Poult. Sci. 84:601-606. Wilson, H.R. 1991. Interrelationships of egg size, chick size, posthatching growth and hatchability. W. Poult. Sci. J. 47:5-20.

76

FEED MANUFACTURING CONSIDERATIONS FOR USING DDGS IN POULTRY AND LIVESTOCK DIETS
Dr. Keith C. Behnke Dept. of Grain Science and Industry Kansas State University 201 Shellenberger Hall Manhattan, KS 66506-2201 Phone: 785-532-4083 FAX: 785-532-4017 Email: kbfeed@ksu.edu Summary While not a new ingredient in livestock feed, the volume available and the relative price of DDGS have forced many feed manufacturers into using greater levels than ever before. The use of DDGS has caused feed manufacturing problems at nearly every phase of feed manufacturing. These include railcars that simply wont unload, feeder screws and supply bins that are wrong for the ingredient, nutrient variation that results in out-of-spec feeds leaving the feed mill, and, of greatest concern, pellet throughput and pellet quality concerns. Nearly all of these problems are the result of the physical properties of DDGS and of the way a particular ethanol facility might manage their byproducts. There is little doubt that the ethanol industry will continue to grow, displacing feed corn with DDGS and other byproducts. It is imperative that we learn to deal with these byproducts effectively so that we can produce the highest quality feeds possible. Introduction Dried distillers grains (DDGS), in one form or another, have been used in compound feeds or as livestock feeds for nearly a century. Early on, of course, the distillers grains came from the beverage alcohol industry and were a very minor ingredient available on a regional basis. Things have certainly changed and everybody is trying to get into the game. What this has meant to the livestock industry is a significant increase in grain prices and in the availability of DDGS. Nutritionists in all livestock areas are scrambling to keep feed costs as low as possible while keeping the desired performance in efficiency and growth rate. Historically, DDGS have been used primarily in ruminant diets. While that is still true today, DDGS are finding their way into poultry and swine diets at an ever-increasing level. As is always the case, the introduction of a previously unused ingredient is not without frustration at most feed mills. The purpose of this paper is to identify some of the issues that have been identified and to discuss ways in which the resulting problems can be addressed. Feed Mill Issues with DDGS Bin Space Allocation Anytime a new ingredient is introduced into a feed mill, the first issue to be dealt with is bin and storage space allocation. It is seldom that the feed mill will have an open bin above the major ingredient scale that can be assigned to the new ingredient. When making bin assignments, the minimum information needed by the manager would include the expected usage rate (pounds per ton), physical properties such as density and flow characteristics, and an estimate of how long the ingredient will be used, in the case of seasonal ingredients.
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

77

In many feed mills, there is simply no way to open a bin for a new ingredient without significantly disrupting production. In these cases, the best solution is to discontinue using an existing ingredient and designate the former ingredient bin for the new ingredient. However, if the bin volume, hopper configuration, and feeder screw design are not compatible for use with the new ingredient, further re-arranging will be necessary. For example, lets assume that DDGS is coming into the feed mill and Phu Phu meal is being discontinued. Phu Phu meal is fairly heavy at 60 lbs/ft3 while DDGS is fairly light at 30 lbs/ft3. Assuming nothing is changed, it will take twice as long for the same weight of DDGS to be delivered as it took for the Phu Phu meal. The obvious question is can the batching cycle be extended without a negative effect on the mixing cycle? Receiving and Storage One would have to live in a cave to not have heard horror stories about receiving and unloading railcars of DDGS. Because of very limited storage at ethanol sites, DDGS are often loaded directly from the dryer onto railcars. In most ethanol plants, the solubles are condensed to about 50% DM, and then sprayed onto the grains just before entering the dryer. While, on average, the moisture content of the DDGS is appropriate (<11%) for shipping and storage, the likely variation within the mass is substantial. Because of the water binding capacity of the solubles, it is likely that they will be at 15-18% moisture while the fermentation solids (fiber, protein, and yeast) are below the 10-11% target. When blended, the internal moisture will equilibrate but not before serious bridging can occur. At this point in time, anti-caking agents, such as those used in soybean meal, are not approved for use in DDGS and there is probably some question as to how well they would work in this application. The real culprit in DDGS causing the arch-formation, in addition to the solubles in DDGS, is the flat, plate-like structure of the bran particles. Soybean meal has a very similar shape as a result of the flaking process prior to solvent extraction. There are several approaches that can improve the flow characteristics of DDGS. Unfortunately, the two most logical must occur at the ethanol site. The first is to simply hold the DDGS in storage onsite until moisture equilibration occurs. In most cases, that would be five to seven days. After equilibration occurs and the matrix is broken, bridging and arch-formation are unlikely to occur in normal transportation systems. The second approach would be to pellet the DDGS (Bauer and Clark, 2004). In this experiment, DDGS were pelleted using various conditioning temperatures and die sizes to determine the ease of pelleting and to evaluate the physical properties and flow characteristics of the resulting product. The results of the study would indicate that nearly any level of agglomeration improved the flowability of DDGS. If this approach is ultimately adopted by the ethanol industry, there will be costs passed on to the ingredient buyer for the pelleting. In addition, if a good quality pellet were produced, grinding at the feed mill would be required thus adding cost to the ingredient. However, there are real costs to the feed mill in labor and receiving system downtime while trying to unload a bridged railcar. In addition, worker safety must be considered as well. Without significant changes at the source of the DDGS, the feed industry will have to find ways, other than the sledgehammer approach, to efficiently unload railcars of DDGS. In several California feed mills, where 20 plus railcars of DDGS arrive at a time, a stationary device is mounted above the rail pit that is used to drive a spear down through the grains to break the bridge. The device is essentially a backhoe equipped with a 10' to 12' pipe spear. While the device is effective at unloading a railcar in a reasonable amount of time, additional labor, capital, and operating expenses are obviously necessary. It appears that, once the moisture is equilibrated and the initial arch-formations are destroyed, DDGS will flow well within the feed mill systems.

78

Variation While nutrient variation is not directly a feed manufacturing problem, it becomes an issue when quality control sample assays come back out of tolerance and the feed mill operators are called on the carpet. In 2004, nutritional scientists with the Degussa Corporation published a summary of assay results for 51 samples of DDGS from various ethanol plants in the U.S. Crude protein (CP) content of these samples ranged from a low of 25.1% CP to a high of 31.1% CP with an average of 27.5% CP and CV of 5.1%. At a use level of 5% (100 lbs/ton), this degree of variability would result in a protein difference of 0.177% CP in the final feed. In todays world, nutritionists are often pushing the level of inclusion to 10 to 12%, and, because of demand, the ingredients are coming from several ethanol plants over a wider geographical area. Both of these factors could result in greater nutrient variation in finished feeds. At the 2006 Mid-Atlantic Nutrition Conference, data presented by Fiene et al. (2006) documented the variability in predicted lysine digestibility in DDGS from eight different sources. The mean predicted digestibilities ranged from 56.9% to a high of 72.2%. From the data, it was obvious that some suppliers were doing a much better job of controlling their processes than others with a high range of nearly 28% (67.5% AV, 27.8% range) and a low range of just 3.9% (72.2% AV, 3.9% range). The obvious question is what can be done at the feed mill level to smooth the inherent nutrient variability? The first answer is based on the implication above in that, if an ingredient is sourced from a single supply point (i.e., ethanol facility in the case of DDGS), the nutrient variation and other quality attributes are usually lower than if multiple suppliers are used. In most cases, feed mill managers have little influence over purchasing decisions. However, if the feed mill is doing a good job of sampling and tracking suppliers, careful analysis of the data may reveal patterns that, when properly presented, can influence purchasing decisions. Dont expect your purchasing department to do the data analyses. They typically focus on one thing only and that is price delivered. Tracking nutrient variation in feed ingredients is not a new concept. Deyoe (1964) published an extensive data set regarding ingredient variation and its effect on animal nutrition. Knowing that variations exist and actually being proactive in doing something about it are two different things. Chung and Pfost (1976) address the issue of overcoming ingredient variation in the feed mill. The two major techniques discussed rely on having a good sense of the existing or potential variation in critical nutrients for a specific ingredient. The useful technique discussed involves blending several lots of an ingredient so that each lot is equally represented in any sample taken. If this can be done, then: 0 = m (the new average for the blend is equal to the average of all lot means) and: s =

s n

where s is the new blend standard deviation, s is the old standard deviation of the

individual lots prior to blending, and n is the number of lots blended. It is obvious that this type of blending can take place only in facilities equipped to do so. For example, a grain terminal, where ingredients can be drawn from numerous storage bins simultaneously, would be ideal. The second procedure is much more practical and involves simply segregating an ingredient into two separate storage facilities based on a lot being above or below a predetermined value (e.g., historical average). By segregating, it can be shown that:

x high
and

= m +

79

xlow

= m

S and S

= 0.603S .

When this is done, it is a simple matter for the nutritionist to formulate rations using the same ingredient with two different nutrient values by treating them as different ingredients. For example, if we segregated corn based on an average of 7% with an S of .75%, then we would have a supply of corn with an average of 7.6% (SD = 0.45%) and a supply of 6.4% (SD = 0.45%). In the above example, the only costs involved are the labor and management involved in inventorying two ingredients instead of one. Of course, a way to quickly analyze for the attribute being used as the basis for segregation is needed, but many feed mills already have an NIR available for rapid testing of moisture, protein, and oil. While the use level of DDGS, at this time, may not be sufficient to justify implementing the above procedure, it is equally useful for grain and protein meals as well and can represent the next level of management skills needed in modern mills. Pelleting and Pellet Quality with DDGS It is likely that the biggest issue in feed manufacturing with DDGS is the effect on pellet throughput and pellet quality. Unfortunately, there are few publications and scientific articles written on this subject and the majority of information is antiquated in nature. It is pretty well accepted that, when the level of DDGS in the formula exceeds 5-7%, pellet throughput, together with pellet quality, will suffer. The questions that need to be addressed are: why does this happen and what can be done to correct the problem? It is unusual that pellet quality is reduced as pelleting rate is decreased. When throughput is reduced and if nothing else changes, pellet quality usually increases. That is due primarily to the increased time a given pellet stays in the die hole where the bonds that hold particles together are formed. It has been documented (Behnke and Beyer, 2002) that starch is usually involved in the bonding between particles that results in strong durable pellets. The fact that there is little starch in DDGS that can be gelatinized and made into an adhesive contributes to poor particle bonding. In addition, DDGS are relatively high in oil compared to the grain and protein meal it replaces. Again, the bonds between particles in the pellet are affected because they are primarily hydrophilic in nature. If sufficient oil is present to coat the particles to some extent, the hydrophobic nature of the coating inhibits bonding between starches, proteins, and the like. As to the throughput issue, this is the feed manufacturing version of the perfect storm. Because of environmental concerns, many feed mills have essentially stopped using dicalcium phosphate (dical) or deflorinated phosphate (deflor) as a mineral in formulas. Instead, the enzyme phytase is included to improve the availability of organic phosphorus. It has long been recognized that either mineral, but particularly deflorinated phosphate, contributes to increased pelleting rate. It is accepted that the minerals contribute an abrasive character to the feed thus polishing the surface of the die holes, therefore, reducing the level of friction between the pellet and the die hole surface. It is theorized that solubilized protein and/or crystalline starch go through what is known as a glass transition and are essentially burned onto the die hole surface. Even though the pellet mash and pellets are about 180-190F, the die itself is likely about 300F which is sufficient to cause the glass transition of proteins and essentially bond the protein to the die hole surface. It is likely that the culprit in DDGS is the corn protein, zein. The processes of grinding, fermentation, and distillation frees this insoluble protein and it certainly is concentrated relative to the level in ground corn. As to the question of what to do about it, there are no easy answers. One might be tempted to use a thinner die which would allow increased throughput. However, there would also be a negative impact

80

on pellet quality which is already an issue. Some feed mills have found it useful to use fine silica sand in some formulas to re-condition the die periodically. If used, the particle size of the sand should be approximately the same as defluorinated phosphate to be effective at scouring the die. The most convenient approach would be to select a formula that is run fairly often (e.g., every two hours) and simply include sand as an ingredient. Using an ingredient, such as sand, in a commercial feed presents a rather special problem due to required labeling. Imagine the reaction of a feed purchaser when reviewing the ingredient list and finding the word sand. One might choose to use the words silica or silicon dioxide instead of sand. As to the future in dealing with pelleting issues related to higher levels of formula DDGS, there are several possibilities. It may be that the die manufacturers can help with more appropriate alloys. There is no doubt that the DDGS themselves will change as the ethanol industry evolves. Several of the newer facilities are being designed with front-end fractionation. This topic will be discussed by another speaker at the Conference so will be only briefly discussed here. Essentially, front-end fractionation involves removing most of the bran and germ prior to grinding and conversion of the starch to sugar. What this will mean is that the DDGS will be substantially higher in protein and lower in oil and fiber. There is no doubt that this will result in changes at the pellet mill that will be as dramatic as when DDGS first found their way into swine and poultry diets. Conclusion While not a new ingredient in livestock feed, the volume available and the relative price of DDGS have forced many feed manufacturers into using greater levels than ever before. The use of DDGS has caused feed manufacturing problems at nearly every phase of feed manufacturing. These include railcars that simply wont unload, feeder screws and supply bins that are wrong for the ingredient, nutrient variation that results in out-of-spec feeds leaving the feed mill, and, of greatest concern, pellet throughput and pellet quality concerns. Nearly all of these problems are the result of the physical properties of DDGS and of the way a particular ethanol facility might manage their byproducts. There is little doubt that the ethanol industry will continue to grow, displacing feed corn with DDGS and other byproducts. It is imperative that we learn to deal with these byproducts effectively so that we can produce the highest quality feeds possible. References Bauer, L., and P.M. Clark. 2004. Increasing the flowability and handling of distillers dried grains with solubles. Unpublished data. Kansas State University. Behnke, K.C., and R. Scott Beyer. 2002. Effect of feed processing on broiler performance. VIII International Seminar on Poultry Production and Pathology. Santiago, Chile. Chung, D.S., and H.B. Pfost. 1976. Overcoming the effects of ingredient variation. Pages 56-58. In: Feed Manufacturing Technology. American Feed Manufacturers Assn., Arlington, VA. Deyoe, C.W. 1964. Variation in the composition of feed ingredients. Feed Age 14(8). Fiene, S.P., T. Walsh-York and C. Schasteen. 2006. Correlation of DDGS IDEA digestibility assay for poultry with cockerel true amino acid digestibility. Pages 82-89. In: Proc. 4th Mid-Atlantic Nutrition Conference. Zimmermann, N. (ed.). College Park, MD 20742.

81

PROCESS AND ENGINEERING EFFECTS ON DDGS PRODUCTS PRESENT AND FUTURE Vijay Singh1*, Carl Parsons2 and Jim Pettigrew2 Agricultural and Biological Engineering Department 2 Animal Science Department University of Illinois at Urbana-Champaign *1304 W. Pennsylvania Ave. Urbana, IL 61801 Phone: 217-333-9510 Email: vsingh@uiuc.edu Summary In a conventional dry grind process, corn is processed to produce ethanol and a low valued animal food coproduct called distillers dried grains with solubles (DDGS). Approximately 33% of corn in dry grind ethanol plant becomes DDGS. Due to its high fiber content DDGS has traditionally being sold as ruminant foodstuffs. New fractionation technologies are being implemented to recovery valuable coproducts, reduce amount of DDGS produced and improve fermentation efficiency in conventional dry grind ethanol plants. These technologies include corn fractionation as well as DDGS fractionation. Corn fractionation can be broadly classified as wet and dry technologies. Wet fractionation involves a short soaking of corn followed by milling to recover germ, pericarp fiber and/or endosperm fiber in an aqueous medium prior to fermentation of degermed defibered slurry. In dry fractionation, a dry degerm defiber process is used to separate germ and pericarp fiber prior to fermentation of the endosperm fraction. Both wet and dry processes reduce the total amount of DDGS produced, increase it protein content and reduce its fiber content. Depending upon the modified process used, the amount of DDGS produced can be reduced by 70% and its protein content can be increased to 58%. DDGS fractionation involves sieving and elutriation (aspiration) to separate fiber from DDGS. This process recovers fiber as a coproduct, increases protein and fat content of residual DDGS and reduces the fiber content of residual DDGS. Depending upon the parameters used this process increased protein and fat contents of residual DDGS from 28 to 41% and 12 to 14%, respectively. A reduction in fiber content and increase in protein content of DDGS could allow increased use of DDGS as nonruminant foodstuffs. Introduction Dry grind ethanol production from corn is growing at fast pace in the US. In last 4 years ethanol production has increased 126% (RFA, 2006). This increase in dry grind ethanol production is expected to continue for next several years and it is estimated to reach 12 billion gallons by 2012. Most of this increase in ethanol production will come from construction of new dry grind ethanol plants. In a conventional dry grind process, corn is ground and mixed with water to produce slurry. The slurry is cooked; slurry starch is liquefied, saccharified and fermented to produce ethanol. The remaining nonfermentables (germ, fiber and protein) are recovered together at the end of the dry grind process as an animal food coproduct called distillers dried grains with solubles (DDGS). With increase in ethanol production, the amount of DDGS will increase concomitantly. DDGS due to its high fiber content is mainly used as foodstuffs in ruminant (dairy and beef cattle) diets and is a low valued coproduct. There is a need to recover valuable coproducts, reduce volume of DDGS and improve its nutritional characteristics for increasing use in non ruminant (poultry and swine) diets.

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

82

Composition of Corn A corn kernel has four main parts: 1) tip cap, 2) pericarp, 3) germ and 4) endosperm. Watson (2003) gave the percent component parts and the composition of these parts of dent corn kernels, as shown in Table 1. There are four kinds of protein in corn kernel based on their solubility. Osborne (1924) classified corn protein as: albumins-proteins soluble in water; globulins-proteins soluble in dilute salt solutions; prolaminsproteins soluble in 70% alcohol solution; and glutelins-proteins soluble in dilute acid or based. Lawton and Wilson (2003) reviewed the corn protein composition for dent corn reported in literature (Table 2). Albumins and globulins are physiologically active protein (enzymes) and are concentrated in germ, aleurone and pericarp fractions. Small amounts of albumins and globulins (5% of total endosperm protein) are found in endosperm fraction (Hoseney 1994). Albumins and globulins have good amino acid balance and are high in Lysine, Tryptophan and Methionine. Prolamins and glutelins are classified as storage proteins and constitute 72% of total endosperm protein. Prolamins and glutelins are deficient in Lysine, Tryptophan and Methionine. Germ comprises of 83% of the fat and 26% of the protein in the corn kernel (Table 2). Most of the phytic acid in corn kernel is in the germ fraction. There are two kinds of fiber in corn kernel: pericarp fiber and endosperm fiber. Pericarp fiber is coarse fiber fraction comprising of dead cell wall material surrounding the corn kernel. Pericarp fiber constitutes 50% of the fiber in the corn kernel. Endosperm fiber is fine fiber fraction comprising of cellular material inside the corn endosperm. Table 1. Whole corn kernel composition and composition of its fractions (endosperm, germ, pericarp and tip cap)1. % (db) Composition of Whole Kernel Starch Fat Protein Ash Sugar Fiber Whole Kernel 73.4 4.4 9.1 1.4 1.9 9.5 Kernel Fractions Percent of Total Indicated Constituents in Kernel Fraction Endosperm 98.1 15.4 73.8 17.9 28.6 27 Germ 1.5 82.6 26.2 78.4 69.3 16 Pericarp 0.6 1.3 2.6 2.9 1.2 51 Tip Cap 0.1 0.8 0.9 1 0.8 0.01 1 Data from Watson (2003). Table 2. Distribution of corn endosperm protein in dent corn1 Corn 1 Corn 2 Corn 3 Albumin 7.8 12.4 7.8 Globulin 0 0 0 Prolamin 50 33.9 37.6 Glutelins 38.2 36.8 43.6 Residue 4 16.9 11 1 Data from Lawton and Wilson (2003). CONVENTIONAL DRY GRIND PROCESS A schematic of the dry grind process is shown in Figure 1. In the conventional dry grind process, the kernel is ground using a hammermill. Dry granular material is mixed with water to form slurry, which is cooked at approximately 160C using pressurized steam to break down the crystalline structure of starch granules. Alpha-amylase is added to break down starch polymers into short chain molecules, called dextrins, to form mash. The mash is held at an elevated temperature (~70C) for a short period of time, cooled to 32C and transferred into a fermentation vessel. Glucoamylase and yeast are added for simultaneous saccharification and fermentation. In the mash, glucoamylase breaks down dextrins into mono or

Corn 4 4.7 3.5 45.8 38 9

83

disaccharides, such as glucose and maltose, while yeast ferment these saccharides into ethanol. At the end of fermentation, the resulting beer is transferred to a holding tank called a beer well. From the beer well, the beer is transferred to a stripper/rectifier column to remove ethanol. Overflow from the stripper/rectifier column is an ethanol and water mixture and underflow from the column is whole stillage (nonfermentable components of corn, yeast and water). The ethanol and water mixture is processed further through a distillation column and molecular sieves to remove remaining water from the ethanol.

Corn

Grinding (Hammermill) Water Mash Overhead product (Recycled back)

Blending Enzymes

CO2

Liquefaction

Yeast & Enzymes Saccharification Stripping/ & Fermentation Rectifying column

Dehydration column

Ethanol
Thin Stillage Syrup

Centrifuge Wet Grains

Evaporator

DDGS
Figure 1. Conventional corn dry grind process. Whole stillage (WS) is centrifuged to produce thin stillage (water and soluble solids) and wet grains (suspended solids). Using an evaporator, thin stillage (TS) is concentrated into syrup and mixed with the wet grains (WG), which is dried to produce a coproduct with 12% moisture content. This coproduct is marketed as DDGS. Modified Dry Grind Corn Processes Wet fractionation of corn prior to fermentation: enzymatic dry grind corn process A modified dry grind process which involves corn fractionation in an aqueous medium to recover germ, pericarp fiber and endosperm fiber as valuable coproducts has been developed (Figure 2) (Singh et al., 2005). This modified dry grind ethanol process is known as the enzymatic dry grind (E-Mill) corn process. The E-Mill process involves soaking corn kernels in water for a short period of time (6 to 12 hr) followed by coarse grinding and incubating with protease and starch degrading enzymes for 2 to 4 hr (Figure 2). Protease and starch degrading enzymes increase specific gravity of the slurry and aid in separation of individual corn components. Germ and pericarp fiber are recovered by floatation (hydrocylcones) (Singh and Eckhoff 1996; Singh et al., 1999; Wahjudi et al., 2001). Endosperm fiber can be recovered by use of screens (200 mesh or 0.074 mm opening) either prior to fermentation (Singh et al., 2005) or after fermentation (Wang et al., 2005). Recovery of endosperm fiber after fermentation reduces the loss of starch in fiber fraction and increases ethanol yield. Rest of the ground corn slurry is processed for ethanol production. E-Mill process benefits dry

84

grind ethanol production in three ways: 1) by adding valuable coproducts (corn germ, pericarp fiber and endosperm fiber) to the process, 2) by increasing the plant capacity and 3) by increasing the amount of protein and reducing the amount of fiber in DDGS. Currently in the US there are two dry grind corn plants using EMill process.
Corn Soaking Incubation Enzymes

Pericarp Fiber Germ & Fiber Dryer Grinding Germ & Fiber (Degermination mill) Germ Air Germ clones Aspirator Fine Grinding (Degermination mill) Enzymes Overhead (Recycled back)

CO 2 Liquefaction

Ethanol Yeast Dehydration column Thin & Enzymes Saccharification Stripping/ Centrifuge Stillage & Fermentation Rectifying Syrup Wet Grains column

Evaporator

QGQF DDGS
Figure 2. Enzymatic (E-Mill) dry grind corn process. Comparison of DDDG from E-Mill and Conventional Dry Grind Processes DDGS protein content was 28 and 58% for conventional and E-Mill processes, respectively (Table 3). Protein content of DDGS for the E-Mill process was higher than protein content of other high protein foodstuffs such as soybean meal (54%). Fat contents of DDGS materials were 12.7 and 4.5% for conventional and E-Mill processes, respectively. No differences were observed in ash contents of DDGS. Due to process modification, DDGS acid detergent fiber (ADF) content was reduced. Compared to conventional DDGS, ADF content was reduced from 10.8 to 2.0% for the E-Mill process (Table 3). E-Mill process reduces the volume of the DDGS by approximately 70%, increases the protein content and reduced the fiber content compared to the conventional dry grind process. Higher protein and lower fiber content can diversify DDGS as a more valuable foodstuff for nonruminant animals. This is important because the predicted growth in ethanol industry could lead to over production of conventional DDGS and limited market demand as ruminant foodstuffs.

85

Table 3. Distiller dried grains with solubles (DDGS) composition of conventional (Conv.) and enzymatic milling (E-Mill) dry grind ethanol processes1. Conv. E-Mill CGM* SBM Crude Protein (%) 28.5 58.5 66.7 53.9 Crude Fat (%) 12.7 4.5 2.8 1.1 Ash (%) 3.6 3.2 --Acid Detergent Fiber (%) 10.8 2.0 6.9 5.9 1 Data form Singh et al. (2005). 2 CGM: corn gluten meal; SBM: soybean meal. Dry Fractionation of Corn Prior to Fermentation: Dry Degerm Defiber Process Another modified dry grind process uses corn dry fractionation to recovery germ and pericarp fiber as valuable coproducts prior to fermentation (Figure 3) (Murthy et al., 2006). This process is called dry degerm defiber (3D) process. In 3D process, corn is tempered with hot water or steam for short period of time (5 to 10 min) and ground in a degerminator to remove germ and pericarp from corn endosperm (Duensing et al., 2003). During grinding corn endosperm is broken into smaller pieces (grits). Germ is separated from grits with the help of gravity tables (density separation) and fiber is removed from grits by aspiration. Grits are further ground to reduce particle size and processed using conventional dry grind ethanol methods to produce ethanol and DDGS. Endosperm fiber is not recovered in 3D process. Currently in the US there are three dry grind corn plants using 3D process.
Stea Pericarp Fiber Tails (Grits) Roller Mill Sifter Sifte Ger

Corn Cor

Beall Degerminato

Through (Germ Pericarp Fiber Endosper Fraction Water Mash

Hammer Mill Saccharification & C 2 Overhead (Recycled back) Dehydration

Liquefactio Yeast & Stripping Rectifyin colum

Ethanol
Centrifug Wet Grains Thi Stillage Syru Evaporato

3D Figure 3. The dry degerm defiber (3D) process.

86

Comparison of DDGS from Wet and Dry Fractionation Processes Martinez-Amezcua (2005) evaluated nutrient composition of DDGS samples produced using laboratory wet and dry fractionation processes. In dry fractionation process (3D process) endosperm fiber is not recovered and proteases are not used in the process. To allow comparison of dry fractionation with wet fractionation process, endosperm fiber recovery and use of protease were eliminated from wet fractionation process. The modified wet fractionation process was called quick germ quick fiber (QGQF process). The DDGS from wet and dry fractionation processes were compared to DDGS produced using laboratory conventional dry grind process (Table 4). Table 4. Distiller dried grains with solubles (DDGS) composition of conventional (Conv.), dry degerm defiber (3D) and quick germ quick fiber (QGQF) dry grind ethanol processes1 Conv. Crude Protein (%) 21.2 Crude Fat (%) 13.9 Fiber (TDF) 36.4 Lysine (%) 0.73 Lys, % of CP 3.4 Total phosphorus (%) 0.78 1 Data from Martinez-Amezcua (2005). 3D 23.8 8.7 28.0 0.63 2.5 0.47 QGQF 28.0 12.6 25.3 0.91 3.3 1.12

Crude protein (CP) of both wet and dry fractionation processes was higher than CP of conventional DDGS. This increase in CP was expected because germ and fiber dilute the protein content in conventional DDGS and their removal will result in higher protein content. Among the two fractionation process, CP of wet process (QGQF process) was higher than dry process (3D process). The higher CP of the QGQF DDGS was possibly due to cleaner separation of germ and fiber from endosperm (less loss of protein) and due to leaching of soluble proteins during the soaking process; the water soluble fraction was used in fermentation process and was concentrated in the final DDGS. Water soluble proteins (albumins and globulins) leach out of germ during soaking in wet fractionation process and get concentrated in DDGS. Whereas in dry fractionation process these protein are lost with the germ fraction and are not recovered in DDGS. That is why the lysine content of QGQF DDGS was higher than 3D or Conventional DDGS. The DDGS produced by the 3D and QGQF processes had lower concentrations of fat than the conventional DDGS. The lower fat was due to the removal of germ. Total dietary fiber was reduced from 36% in the conventional DDGS sample to 28 and 25% by the 3D and QGQF methods. The P content of DDGS was reduced by the 3D process but was increased by the QGQF process. A reduction for 3D was expected since much of the germ is removed and almost 90% of the phytic acid is present in the germ of corn (Ravindran et al., 1995; Rebollar and Mateos, 1999). The increase in P for the QGQF was unexpected and may have been due to leaching of P during the 12 hr soaking process. Removal of Fiber from DDGS: Elusieve Process A process called elusieve has been developed to separate fiber from distillers dried grains with solubles (DDGS). Separation of fiber from DDGS in a dry grind ethanol plant increases protein and fat content and reduces fiber content in the resulting DDGS. Fiber produced from the elusive process can be used for recovery of other value added coproducts. The elusieve process uses sieving and elutriation to separate fiber from DDGS (Figure 4).

87

Corn

Grinding (Hammermill) Water Mash Blending Enzymes CO 2 Overhead product (Recycled back)

Liquefaction Dehydration Ethanol Yeast column & Enzymes Thin Centrifuge Saccharification Stillage Stripping/ Evaporator & Fermentation Rectifying Wet Grains Syrup column

DDGS

24 T 34 T 35 M 60 M Pan

H = Heavier Fraction L = Lighter Fraction


L H H H H

L L L

Enhanced DDGS

Fiber

Figure 4. Elusieve process to remove fiber from DDGS. Material carried to the top of the elutriation column is called lighter fraction or fiber fraction and material that settled to the bottom of the column is called heavier fraction or enhanced DDGS. Conventional DDGS samples, obtained from dry grind corn plants, were processed using elusive technology. By adjusting process parameters, elusive processing increased protein and fat contents of enhanced DDGS from 28 to 41% and 12 to 14%, respectively, and reduced neutral detergent fiber content from 32 to 19%, compared to the original DDGS (Table 5 and 6.) (Srinivasan et al., 2005). Elusive process is low cost solution to the reduce fiber content of conventional DDGS. The payback period for elusieve process for a dry grind ethanol plant producing 40 million gallons per year was estimated to be less than two years (Srinivasan et al., 2006). Table 5. Composition of different size materials after sieving of commercial DDGS sample1. Nominal % (w/w) Protein Particle Size Retained (%) (Microns) on Screen Original DDGS All 100 33.6 Material on 24T2 >869 27 29.3 Material on 34T 582 to 869 19.4 26.9 Material on 35M 447 to 582 13.3 31.2 Material on 60M 234 to 447 20.1 37.5 Material in Pan <234 20.2 42.2 1 Data from Srinivasan et al. (2005). 2 Screen size, M and T refer to market grade cloth and tensil bolt cloth. Size Category Fat (%) 12.5 12.5 11.3 10.9 11.3 12.9 Neutral Detergent Fiber (%) 32.5 33.4 37.8 33.6 29.3 19.0

88

Table 6. Elutriation (aspiration) of fiber from material on 24T screen1. Fraction Neutral Detergent Fiber (%) Lighter Fraction 53.3 Material on 24T 33.4 Enhanced DDGS 32.6 1 Data from Srinivasan et al. (2005). Conclusions Modified dry grind processes have been developed that involve fractionation of corn at the beginning of the dry grind process and recovery of nonfermentable components (germ, pericarp and endosperm fiber) prior to fermentation. Other modified processes involve fractionation of conventional DDGS recover fiber as a coproduct. These technologies reduce the amount of DDGS produced in a dry grind ethanol plant and improve its nutritional composition. References Hoseney, R.R. 1994. Proteins of cereals. Pages 65-79. In: Principles of Cereal Sciences and Technology, 2nd ed. Am. Assoc. Cereal Chem., St. Paul, MN. Lawton, J.W., and C.M. Wilson. 2003. Protein of kernel. Pages 313-354. In: Corn: Chemistry and Technology. 2nd edition. White, P.J., and L.A. Johnson (eds.). American Association of Cereal Chemists, St. Paul, MN. Maisch, W.F. 2003. Fermentation processes and products. Pages 695-721. In: Corn: Chemistry and Technology. 2nd edition. White, P.J., and L.A. Johnson (eds.). American Association of Cereal Chemists, St. Paul, MN. Martinez-Amezcua, C. 2005. Nutritional evaluation of corn distillers dried grains with solubles for poultry. PhD thesis. University of Illinois at Urbana-Champaign. Murthy, G.S., V. Singh, D.B. Johnston, K.D. Rausch and M.E. Tumbleson. 2006. Evaluation and strategies to improve fermentation characteristics of modified dry grind corn processes. Cereal Chem. 83:455-459. Osborne, T.B. 1924. The vegetable proteins. 2nd ed. Longmans, Green and Co., London, UK. Ravindran, V., W.L. Bryden and E.T. Kornegay. 1995. Phytates: Occurrence, bioavailability and implications in poultry nutrition. Poult. Avian Biol. Rev. 6:125-143. Rebollar, P.G., and G.G. Mateos. 1999. El fosforo en nutricion animal. Necesidades, valoracion de materias primas y mejora de la disponibilidad. Pages 19-64. In: XV Curso de especializacion. Avances en Nutricion y Alimentacion Animal. Organized by FEDNA. Madrid, Spain. RFA. 2006. Ethanol industry outlook 2006. Washington, D.C. Renewable Fuels Association. Available at: www.ethanolrfa.org/objects/pdf/outlook/outlook_2006.pdf. Accessed 1 February 2007. Singh, V., and S.R. Eckhoff. 1996. Effect of soak time, soak temperature and lactic acid on germ recovery parameters. Cereal Chem. 73:716-720. Singh, V., D.B. Johnston, K. Naidu, K.D. Rausch, R.L. Belyea and M.E. Tumbleson. 2005. Comparison of modified dry grind corn processes for fermentation characteristics and DDGS composition. Cereal Chem. 82:187-190. Singh, V., R.A. Moreau, L.W. Doner, S.R. Eckhoff and K.B. Hicks. 1999. Recovery of fiber in the corn drygrind ethanol process: a feedstock for valuable coproducts. Cereal Chem. 76:868-872. Srinivasan, R., R.A. Moreau, K.D. Rausch, R.L. Belyea, M.E. Tumbleson and V. Singh. 2005. Separation of fiber from distillers dried grains with solubles (ddgs) using sieving and elutriation. Cereal Chem. 82:528533. Protein (%) 19.3 29.3 35.6 Fat (%) 7.05 12.5 14.2

89

Srinivasan, R., V. Singh, R.L. Belyea, K.D. Rausch, R.A. Moreau and M.E. Tumbleson. 2006. Economics of fiber separation from distillers dried grains with solubles (DDGS) using sieving and elutriation. Cereal Chem. 83:324-330. Wahjudi, J., L. Xu, P. Wang, V. Singh, P. Buriak, K.D. Rausch, A.J. McAloon, M.E. Tumbleson and S.R. Eckhoff. 2000. Quick fiber process: effect of mash temperature, dry solids and residual germ on fiber yield and purity. Cereal Chem. 77:640-644. Wang, P., V. Singh, L. Xu, D.B. Johnston, K.D. Rausch and M.E. Tumbleson. 2005. Comparison of enzymatic (E-mill) and conventional dry grind corn processes using a granular starch hydrolyzing enzyme. Cereal Chem. 82:734-738. Watson, S.A. 2003. Description, development, structure and composition of the corn kernel. Pages 69-106. In: Corn: Chemistry and Technology. 2nd edition. White, P.J., and L.A. Johnson (eds.). American Association of Cereal Chemists, St. Paul, MN.

90

FORMULATING POULTRY DIETS WITH DDGS HOW FAR CAN WE GO? Sally Noll University of Minnesota Extension Service 1364 Eckles Ave St. Paul, MN 55108 Phone: 612-624-4928 FAX: 612-625-5789 Email: nollx001@umn.edu Co-Authors Carl Parsons1 and William Dozier, III 2 1 Department of Animal Sciences, University of Illinois 2 USDA/ARS, Mississippi State Summary Current feeding trials have examined the use of low and moderate levels of distillers dried grains with solubles (DDGS) inclusion in broiler and turkey diets. In broilers, up to 15% DDGS in grow/finish diets is possible. In market tom turkeys, up to 20% DDGS in grow/finish diets is possible in diets with normal protein content and under conditions where feed intake is maximized. Variability of nutrient content is of concern as risk increases with higher inclusion rates but this variation can be reduced somewhat by using a limited number of sources to provide the material. Some of the nutrient variability in DDGS may be due to addition of different levels of solubles to the wet grains prior to drying. Varying the addition of the solubles to the grains affected particle size, color, and content of fat and minerals. Use of high levels of DDGS will change the amino acid and mineral nutrient profile as well as the amounts of ingredients being used. Introduction Expansion of the ethanol industry in the late 1990s brought about increased supplies of corn DDGS (distillers dried grains with solubles) sparking interest by feed companies and nutritionists to examine the use of this corn co-product in poultry diets. Although DDGS is not a new product, research on feeding of DDGS generated from recently built ethanol plants was not available. Because of cost, availability or quality concerns, the product was not used at all in poultry diets or the use levels were kept conservative with levels above 5% considered high. In the last six months, however, the increase in price of both corn and soybean meal has re-kindled interest in exploring utilization of higher levels of DDGS (levels in excess of 10%). Concerns regarding the use of DDGS still remain quite often that of variability in nutritional and physical characteristics, nutrient quality of the product, and levels of use in poultry diets. This paper will explore the considerations of using higher levels of DDGS in poultry diets based on its nutrient characteristics and the response of poultry to the feeding of high levels of DDGS from published studies. Feeding Trials with DDGS in Meat Type Poultry Chicken Broilers Waldroup et al. (1981) fed up to 25% DDGS in two sets of diets where energy level was allowed to decline (variable) and the other set where diet ME was kept constant at 3200 kcal/kg with the use of supplemental fat (fixed). Diets were formulated without supplemental lysine using NRC ingredient
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

91

specifications in corn-soybean meal based diets and diets were fed as mash. The ratio of nutrients to dietary energy in the variable regimen was kept constant. Diets were fed during 0-21 da and 21-42 da of age. In the diets of fixed energy, BW was not affected by DDGS inclusion level nor was feed efficiency during 0-42 days. In the variable dietary energy regimen, BW was reduced when DDGS was included at levels of 15% or greater. Feed efficiency (gain:feed) was reduced likewise when inclusion levels were 15% or greater. The authors also measured feed density (g/liter). Diets containing 25% DDGS were less dense with density decreasing by 6.9% and 5.4% in the starter and grower diet, respectively. The results of this study indicated that up to 25% of DDGS could be fed with adjustments to the diet energy level. However, broiler growth has changed considerably since this publication and diets were formulated to meet the lysine requirement with intact protein which may have prevented other amino acids from limiting growth. More recent trials were conducted by Lumpkins et al. (2004). Inclusion level of 15% DDGS was examined in diets of low and high-density starter diets. Another trial examined levels of 0, 6, 12, and 18% in diets formulated to be isocaloric and isonitrogenous in starter, grower, and finisher diets. Diets were formulated on a total amino acid basis and supplements of lysine and methionine were used. The base diet was primarily composed of corn-soybean meal with poultry fat used to obtain the desired diet ME level. The DDGS product used was derived from corn used in ethanol production. The product contained 27% crude protein, 9.8 % fat, and total lysine of .85% (as fed basis). Lysine digestibility was determined to be 75% indicating it to be a high quality product. No negative effects on performance were noted by inclusion of DDGS (15%) in the high density starter diets. In the low density diets, feed efficiency (gain:feed) was reduced early (7 and 14 da) when fed diets with 15% DDGS. In the second trial, the feeding of 18% DDGS was found to decrease BW at 16 da of age, an effect which carried over to BW at 42 da of age. Feed efficiency was also worsened during 0-16 da at the high level of DDGS inclusion (18%). No treatment differences were observed during 17-31 da or 0-42 da for feed efficiency. The authors concluded that an 18% level of inclusion was too high for use in starter diets and that a marginal lysine deficiency may have caused the decrease in performance. Market Turkeys In market hens, Roberson (2003) found that level of inclusion and matrix values for DDGS (corn derived) influenced performance of grow/finish market hens. In the first study, four levels of DDGS were fed (0, 9, 18, and 27%). DDGS was incorporated into corn-soybean based diets and supplements of lysine, methionine, and threonine were used. Diets were formulated on a digestible amino acid basis and an energy value of 2870 kcal/kg was used for DDGS ME. In the second trial, diets were formulated on a total amino acid basis to meet 110% of the NRC (1994) amino acid requirements and an energy value of 2805 kcal/kg was assigned to the DDGS. Inclusion levels were 7 and 10% of the diet. In the first study, BW at 105 da was decreased linearly with DDGS inclusion. Feed/gain tended to increase with DDGS inclusion (P<.10). A higher incidence of pendulous crops and greater litter moisture was noted with the 27% inclusion level as compared to diets with 0 or 9% DDGS. In the second trial, growth and feed efficiency were similar among treatments. The acceptable performance with DDGS in the second trial was attributed to the use of higher amino acid specifications to overcome concerns with lysine digestibility and the lowering of the energy level of DDGS from 2870 to 2805 kcal/kg. In market turkeys, previous studies have indicated that up to 10% DDGS could be incorporated into corn-soy diets containing moderate amounts of poultry byproduct meal (PBM) in grow/finish diets for heavy toms and formulated using digestible amino acids (Noll et al., 2002, 2003) . Diets were formulated on a digestible amino acid basis using AME energy values of 2810-2860 kcal/kg for DDGS. In the three studies, comparing the DDGS feeding regimen with the control diet indicated no difference in live performance relative to market body weight and feed conversion (Table 1).

92

Table 1. Corn DDGS and performance of heavy tom turkeys. Experiment Age Season PBM Treatment DDGS BW, Feed/Gain Period Level Level 19 wks Cumulative for study period (wks) (%) (%) (lbs) 1. Noll et al., 5-19 2002 Winter 8-5 Control DDGS 2. Noll et al., 8-19 2003 Winter 8-6 Control DDGS 3. Noll, 2003 11-19 Spring 7-6 Control DDGS 0 12-8 0 11-8 0 10 41.6 41.9 42.4 42.6 40.5 40.2 2.44 2.48 2.62 2.64 2.67 2.63

Lysine digestibility coefficients for DDGS for Exp. 1, 2, and 3 were 78, 64, and 64%, respectively.

As prices or supplies permit, higher levels may be appropriate assuming performance doesnt suffer. Feeding levels of 15 and 20% resulted in performance similar to the control (Table 2) (Noll et al., 2004). However, in a subsequent feeding trial, conducted such that the trial finished in early summer, growth of turkeys fed 20% DDGS was decreased in comparison to the control (Noll et al., 2005). Table 2. High levels of corn DDGS and performance of heavy tom turkeys. Experiment Age Season PBM Treatment DDGS BW, Feed/Gain Period Level Level 19 wks Cumulative for study period (wks) (%) (%) (lbs) 4. Noll et al., 8-19 Winter 7-5 Control 0 38.6 2.80 2004 DDGS 10 38.9 2.80 DDGS 15 39.0 2.80 DDGS 20 38.7 2.80 5. Noll et al., 6-19 2005
1

Summer 8-5

Control DDGS DDGS

0 10 201

38.5 38.3 37.7

2.53 2.52 2.55

Contrast testing indicated the 20% inclusion level to be different from the control (P<.05). Lysine digestibility coefficient for DDGS for Exp. 4 and 5 was 72%.

The difference in the results of feeding higher levels of DDGS between the two trials was attributed to season and protein level of the base diets. The one trial was conducted during the winter where feed intakes would be maximized under cool rearing conditions. Supplemental threonine was not used. The other trial was conducted under summer rearing conditions where feed intake was limiting and supplemental threonine was used to reduce diet crude protein levels.

93

Dietary Considerations in Feeding High Levels of Corn DDGS DDGS Characteristics Several studies have indicated that variability in composition and quality exists in DDGS. When high levels of inclusion are being used, the risk associated with nutrient variability becomes greater. Variability in composition was found in several nutrients. In a study conducted at the University of Minnesota (Ergul et al., 2003; Noll et al., 2003), samples of corn DDGS were collected to determine their nutrient composition and variation among and within sources. Samples (N=20) were obtained from five commercial ethanol plants. Means (as fed basis) for ash, DM, fat, fiber, protein, starch, and sugars were 4.0, 88.3, 10, 5.7, 27.6, 4.7, and 2.3%, respectively. Sources varied in fat, protein, and ash content (P < 0.01). Amino acid content also differed among sources with the exception of Ser (P < 0.05). Respective means for methionine, cystine, lysine, arginine, tryptophan, valine, threonine, and isoleucine were 0.49, 0.52, 0.74, 1.08, 0.22, 1.32, 0.98, and 0.96%. Lysine content was the most variable across all samples (CV=11.2%) followed by cystine (CV=11.3%) and tryptophan (CV=11.1%). Within source, the CV for lysine averaged 4.6%. Respective means for Mg, Na, P, K, Cl, S, and Ca were 0.31, 0.11, 0.73, 0.95, 0.17, 0.65, and 0.03%. Mineral content also varied among sources, with sodium being the most variable among all samples (CV=33%) and was highly variable within sources. Sources of DDGS differed (P < 0.05) in digestibility coefficients for lysine, cystine, threonine, and arginine (71, 77, 72, and 93%, respectively). Sources differed (P < 0.05) in true digestible essential amino acid content except tryptophan. A wide range (0.38 to 0.65%) existed especially for true digestible lysine. Analyses of DDGS indicated that differences in composition are related to source of production during the time period of sample collection. However, within source, composition was found to be relatively consistent with the exception of sodium content. Batal and Dale (2003) also found a large range in sodium content over 12 samples of DDGS (.09.44%, average value .23%). For most nutrients then, obtaining material from one source should help minimize variation among batches of materials. Because lysine is a first or second limiting amino acid in poultry diets and because of its susceptibility to heat damage during the drying process, lysine digestibility will be a major concern in use of DDGS. Other studies have also reported variation among sources in lysine content and digestibility. Table 3 summarizes information from three different studies regarding true lysine digestibility as determined in cecectomized roosters. On average, true digestibility for lysine was in excess of 70% but some individual samples showed low digestibility. If one would take an extreme example where digestible lysine content of DDGS ranged between .39 to .65%, a 20% inclusion level of DDGS would result in a difference of .06% dietary lysine. Table 3. Lysine content and digestibility of DDGS. Source No. of Mean Lysine Mean Lysine Digestibility Samples Content (%) Coefficient (%) Average Range Average Range 1 Ergul et al., 2003 20 .74 .59-.89 71 59-84 Batal and Dale, 20062 8 .71 .39-.86 70 46-76 Fastinger et al., 20061 5 .64 .48-.75 76 65-82
1

As fed basis.

Adjusted to 86% DM.

TMEn was also evaluated in these studies. On an 86% DM basis, TMEn ranged from 2490 to 3190 kcal/kg with an average of 2820 for 17 samples (Batal and Dale, 2006) as determined with conventional roosters. The authors attempted to develop a predictive equation for TMEn based on fat, fiber, protein, and ash content. Fat content was the best predictor of TMEn content, but the overall R2 was quite low (R2=.29). Adding fiber, protein, and ash to the regression model moderately improved the prediction equation (R2=.45). In the study by Fastinger et al. (2006), on an as fed basis, TMEn ranged from 2484 to 3014 kcal/kg with an average of 2871 for the five samples. In this study, the sample with the

94

lowest energy value was associated with the sample having the poorest amino acid digestibility as well. As reported by Abe (2005), TMEn as determined with young turkeys was not affected by source and averaged 2833 kcal/kg agreeing with the value obtained by Batal and Dale (2006). A recent trial tested the assignment of a metabolizable energy value to DDGS (Noll et al., 2005). A grow-finish trial was conducted with turkeys to confirm the appropriate energy value of DDGS to use in diet formulation. Commercial male turkeys (Large White, Hybrid strain) were fed diets varying in level of DDGS (10 or 20% DDGS) and formulated using different levels of MEn assigned to the DDGS during 6 to 19 wks of age. The ME assignments were (kcal/kg): previously determined TMEn in young growing turkeys of 2980; previously determined AMEn with young turkey poults of 2760; and, the NRC (1994) book value of 2480. The basal diet was composed of primarily corn, soybean meal, poultry byproduct meal and .05% supplemental threonine. Diets were formulated on a digestible amino acid basis. A control diet with no DDGS was included. Diets varying in ME assignment did not affect turkey body weight. When the TMEn value was used in formulation, cumulative 6-19 wk f/g was poorer as compared to the NRC value (2.56 vs. 2.52) (P<.05). Determination of energy by TMEn resulted in an overestimation of the energy value of the DDGS when using feed efficiency as the response criteria. While there was no difference in response for the NRC or AMEn energy value, use of the lower NRC energy value could have a large effect on diet cost. Corn derived DDGS can be an economic source of available phosphorus (P). Previous studies have indicated the phosphorus availability of DDGS to be greater than that of phosphorus in corn, the availability of which is estimated at 28% (NRC, 1994). Lumpkins and Batal (2005) obtained P bioavailability estimates of 54 and 68% with chicks. Martinez-Amezcua et al. (2004) found P bioavailability was related to heat processing such that P availability increased from 75 to 87% for a sample of DDGS that was autoclaved. Bioavailability of phosphorus in three other sets of DDGS was 75, 82 and 102%. In a follow-up study, Martinez Amezcua and Parsons (2007) demonstrated that heating or autoclaving DDGS increased P bioavailability; however, digestibility of lysine was decreased. Kalbfleisch and Roberson (2004, 2005) found relatively high availabilities for P, in excess of 85% for DDGS using a turkey poult bioassay. Martinez-Amezcua et al. (2006) found additions of phytase and citric acid in a diet containing 40% DDGS released additional P, improving P availability of the DDGS from 62 to 72%. While the high bioavailability of P could reduce overall diet P content, the large range in bioavailability prevents accurate assignment of P bioavailability. Many things can contribute to this source of variability, such as corn composition, solubles addition, and drying conditions. Variable solubles addition to the wet grains prior to drying could effect the nutrient composition of the dried product and perhaps change the dynamics of the drying process to affect product quality. To examine the effect of the solubles addition, a pilot study was conducted in cooperation with an ethanol plant in Minnesota (Noll et al., 2007). Batches of corn distiller dried grains were produced with varying levels of solubles (syrup) added back to the wet grains (mash). The batches produced contained syrup added at approximately 0, 30, 60, and 100% of the maximum possible addition of syrup to mash. Actual rates of syrup addition were 0, 12, 25, and 42 gal/minute. The different combinations of mash and syrup were dried at the plant. Drying temperature decreased with the decrease in rate of solubles of addition. Digestible amino acids were determined in cecectomized roosters and true metabolizable energy (TMEn) in intact young turkeys. Regression analyses and correlation coefficients (Pearson) were conducted to determine the extent of the relationship between the level of solubles added and the resulting nutrient content. Particle size was greatly affected with larger and more variable particle size observed with the highest level of solubles addition. The larger particles (syrup balls) were readily apparent in the 100% batch. Content of fat and ash increased with solubles addition (Table 4). The TMEn content increased with solubles addition. Mineral content, especially for magnesium, sodium, phosphorus, potassium, chloride, and sulfur increased as the level of solubles addition increased. Protein and amino acid content showed very little change in the various products. True amino acid digestibility coefficients of the essential amino acids tended to be negatively correlated with solubles addition. The results indicate that solubles addition has the largest effect on particle size, color, and contents of fat (and thus TMEn) and minerals.

95

Table 4. Solubles addition and characteristics of the resulting DDGS. Solubles Addition (gal/min) Statistics 42 25 12 0 Correlation P value (Pearson) with solubles addition Product Characteristics Color (CIE Scale) L* 46.1 52.5 56.8 59.4 -.98 .0001 a* 8.8 9.3 8.4 8 .62 .03 b* 35.6 40.4 42.1 43.3 -.92 .0001 Moisture (%) Nutrient (%, DM basis) Protein Fat Fiber Ash TMEn, kcal/kg P, ppm 13.8 10.7 9.75 9.52 .93 .06

32.0 32.5 32.6 32.0 10.5 9.22 9.14 7.97 6.5 10.08 7.76 9.17 4.62 3.72 3.58 2.58 3743 3002 2897 2712 9116 7669 6636 5315

.03 .96 -.51 .97 .94 .99

NS .04 NS .03 .06 .002

DDGS and Diet Composition Incorporation of high levels of DDGS into market turkey and broiler diets can result in potential excesses and deficiencies of several nutrients. Broiler and turkey grower diets formulated with different levels of DDGS show the same trends (Table 5). Use of corn, soybean meal, dicalcium phosphate, and DL-methionine are decreased while supplemental lysine, fat, and calcium carbonate are increased as inclusion level of DDGS increases. Mineral content also changes with decreases in potassium and increases in sulfur content. Protein content increased slightly although changes would be greater if diets were formulated on an amino acid basis. Changes in sodium and chloride content can occur. With phosphorus, in the broiler diets, 20% DDGS resulted in no dicalcium phosphate use. If higher use levels of animal by product are desired, phosphorus levels will become excessive or the amount of DDGS would need to be decreased. The replacement of soybean meal protein with corn protein was associated with declines (total amino acid basis) in tryptophan, arginine, and isoleucine. Content of leucine and valine increased. Production of pelleted feed containing DDGS may have negative effects on feed mill performance. Koch (2006, 2007) presented information on energy usage and pellet quality when adding DDGS in combination with other ingredients. Production of pellets from mixtures of Durum wheat and DDGS resulted in increased energy usage and decreased pellet quality (PDI) with increasing additions of up to 50% DDGS with wheat. Pelleting of a swine diet containing 10% DDGS decreased energy cost but also decreased pellet quality in comparison to pelleting the diet without DDGS.

96

Table 5. Inclusion of DDGS and diet composition. Turkey Grower Diets, 8-11 wks of age1 DDGS Level, % 0 20 30 40 54.65 45.24 40.54 35.83 33.77 23.22 17.95 12.67 4.00 4.00 4.00 4.00 0.00 20.00 30.00 40.00 1.25 0.73 0.47 0.22 0.83 1.20 1.38 1.57 0.17 0.13 0.11 0.09 0.06 0.23 0.31 0.39 0.00 0.00 0.00 0.00 4.53 4.69 4.77 4.84 ++ ++ ++ ++ 22.06 3150 0.82 1.28 1.42 0.23 1.01 0.79 1.74 0.90 22.13 3150 0.82 1.28 1.30 0.20 1.03 0.79 1.94 0.88 22.16 3150 0.82 1.28 1.25 0.18 1.04 0.79 2.04 0.86 22.19 3150 0.82 1.28 1.19 0.16 1.05 0.79 2.14 0.85 Broiler Grower Diets2 DDGS Level, % 0 10 20 63.14 56.19 51.26 27.52 24.28 19.04 5.00 5.00 5.00 0.00 10.00 20.00 0.29 0.02 0.00 0.47 0.64 0.69 0.23 0.22 0.20 0.00 0.01 0.10 0.03 0.00 0.00 2.58 3.01 3.17 ++ ++ ++ 20.44 3150 0.87 1.10 1.29 0.21 0.92 0.75 1.32 0.81 21.14 3150 0.87 1.10 1.29 0.20 0.97 0.75 1.77 0.83 21.17 3150 0.87 1.10 1.23 0.18 0.98 0.75 1.88 0.82

Ingredient (%) Corn Soybean meal Poultry byproduct DDGS Dical. Phosphate Ca. Carbonate Dl met .99 L-Lys HCl Thr Animal fat Other Nutrient (%) Protein ME, kcal/kg Met + Cys Lys Arg Tryp Val Thr Leu Iso
1 2

Formulated based on total amino acid requirements (NRC, 1994) adjusted for 3 wk periods. Formulated based on total amino acid specifications-medium density (Kidd et al ., 2004).

The use of high levels of both animal byproduct and DDGS could replace a considerable quantity of soybean meal protein. A trial was conducted to examine different inclusion levels of poultry byproduct meal (PBM) and DDGS and their combined effect on market tom performance during 5-19 wks of age (Noll et al., 2006). Large White male turkey poults (Nicholas strain) were randomly assigned to pens (10/pen) at 5 wks age and fed one of the following diet treatments (T): 1. Corn and soybean meal control; 2. As T1 with PBM (8% ); 3. As T1 with PBM (12%); 4. As T1 with DDGS (10%); 5. As T1 with DDGS (20%); 6. As T 2 and T4; 7. As T2 and T5; 8. As T3 and T4; and, 9. As T3 and T5. Diets were formulated using digestible amino acids. Diet protein level was established by using intact protein to meet the digestible NRC threonine at 100% of the NRC recommendation. All diets were supplemented as needed with lysine and methionine to meet the specific NRC recommendations for these amino acids. The ratio of calcium:phosphorus was maintained at 2:1 to accommodate the higher levels of phosphorus in the diets containing high levels of PBM and DDGS. Each diet was fed to 10 replicate pens. The experimental design was a completely randomized block design with a factorial arrangement of PBM and DDGS inclusion levels. At 19 wks of age (Table 4), dietary treatment significantly affected 19-wk body weight and feed efficiency (5-19 wks) (P<.001).

97

Table 6. Market tom performance and inclusion of alternative protein sources.

Trt # Diet Description 1 Corn-Soy Control 2 As 1 + 8% PBM 3 As 1 + 12% PBM 4 As 1 + 10% DDGS 5 As 1 + 20% DDGS 6 As Trt 2 & 4 7 As Trt 2 & 5 8 As Trt 3 & 4 9 As Trt 3 & 5 Statistics Treatment P value Treatment LSD (P<.05)

Body Weight (lbs) 11 wks 19 wks 19.12 18.58 18.35 19.20 18.91 18.58 17.87 18.06 18.06
a b bc a a b d cd cd

Feed/Gain 5-19 wks


a ab abc a a ab c abc bc

44.49 43.95 43.77 44.47 44.41 43.99 42.85 43.56 43.09

2.50 2.47 2.46 2.51 2.51 2.50 2.56 2.45 2.55

cd cd d bc c c a d ab

0.0001 0.34

0.0001 0.96

0.0001 0.05

Diets containing PBM (8 or 12%) or DDGS (10 or 20%) were not significantly different from the control. BW of turkeys fed diets containing PBM (8 or 12%) in combination with 20% DDGS was less than that of the control by 3.3%. A significant interaction existed for inclusion of PBM and DDGS (P<.02) for feed efficiency. Feed/gain of turkeys fed diets containing PBM (8 or 12%) or DDGS (10 or 20%) were not significantly different from the control. However, the feed/gain increased for turkeys fed diets containing PBM (8 or 12%) in combination with 20% DDGS and were significantly different from the control by 5 to 6 points. The decrease in performance was suspected to be due to an amino acid deficiency or an excess of the calcium and phosphorus in the diets. Formulating diets with high inclusion levels of DDGS will need to consider potential excesses and deficiencies of certain amino acids and minerals. Phosphorus content of DDGS, while of good bioavailability, may limit its use particularly in combination with animal by product meals. Higher levels of use may result in unacceptable pelleting conditions and high supplemental fat levels. Acknowledgments Technical support - UM Support Staff Jeanine Brannon, Fred Hrbek, Terrance Yourchuck; and, MTGA Nutrition Subcommittee (Dick Nelson, Gary Johnson, Jim Halvorson, Greg Engelke, and George Speers); University of Minnesota research supported in part by a USDA-CSREES Special Research Grant to the Midwest Poultry Consortium, Minnesota Turkey Promotion and Research Council, Minnesota Corn Research Council, DakotaGold Research, ADM, and CSC. References Abe, C. 2005. Distillers Dried Grain with Solubles as a Feed Ingredient for Turkeys. M.S. Thesis, University of Minnesota. Batal, A., and N. Dale. 2003. Mineral composition of distillers dried grains with solubles. J. Appl. Poult. Res. 12:400-403.

98

Batal, A.B., and N.M. Dale. 2006. True metabolizable energy and amino acid digestibility of distillers dried grains with solubles. J. Appl. Poult. Res. 15:89-93. Ergul, T., C. Martinez-Amezcus, C.M. Parsons, B. Walters, J. Brannon and S.L. Noll. 2003. Amino acid digestibility in corn distillers dried grains with solubles. Poult. Sci. 82 (Suppl. 1): 70. Fastinger, N.D., J.D. Latshaw and D.C. Mahan. 2006. Amino acid availability and true metabolizable energy content of corn distillers dried grains with solubles in adult cecectomized roosters. Poult. Sci. 85:1212-1216. Kalbfleisch, J.L., and K.D. Roberson. 2004. Phosphorus availability of distillers dried grains plus solubles for male turkey poults. Poult. Sci. 83 (Suppl. 1):319. Kalbfleisch, J.L., and K.D. Roberson. 2005. Phosphorus availability of distillers dried grains with solubles: Variation in color. Poult. Sci. 84 (Suppl. 1):68. Kidd, M.T., C.D. McDaniel, S.L. Branton, E.R. Miller, B.B. Bore and B.I. Fancher. 2004. Increasing amino acid density improves live performance and carcass yields of commercial broilers. J. Appl. Poult. Res. 13:593-604. Koch, K. 2006. Feed manufacturing with DDGS. Website address: http://www.ddgs.umn.edu/ppt-procstorage-quality.htm. Accessed 2/20/07. Koch, K. 2007. Pelleting and distillers dried grains with solubles. Website address: http://www.ddgs.umn.edu/ppt-proc-storage-quality.htm. Accessed 2/20/07. Lumpkins, B., and A. Batal. 2005. Bioavailability of lysine and phosphorus in distillers dried grains with solubles. Poult. Sci. 84:581-586. Lumpkins, B.S., A.B. Batal and N.M. Dale. 2004. Evaluation of distillers dried grains with solubles as a feed ingredient for broilers. Poult. Sci. 83:1891-1896. Martinez Amezcua, C., C.M. Parsons and D.H. Baker. 2006. Effect of microbial phytase and citric acid on phosphorus bioavailability, apparent metabolizable energy, and amino acid digestibility in distillers dried grains with solubles in chicks. Poult. Sci. 85:470-475. Martinez Amezcua, C., C.M. Parsons and S.L. Noll. 2004. Content and relative bioavailability of phosphorus in distillers dried grains with solubles in chicks. Poult. Sci. 83:971-976. Martinez Amezcua, C., and C.M. Parsons. 2007. Effect of increased heat processing and particle size on phosphorus bioavailability in corn distillers dried grains with solubles. Poult. Sci. 86:331-337. Noll, S.L., C. Abe and J. Brannon. 2003. Nutrient composition of corn distiller dried grains with solubles. Poult. Sci. 82 (Suppl. 1):71. Noll, S.L., C.M. Parsons and J. Brannon. 2007. Nutritional value of corn distillers dried grains with solubles: Influence of solubles addition. Poult. Sci. Submitted Noll, S.L., V. Stangeland, G. Speers, C. Parsons and J. Brannon. 2002. Utilization of canola meal and distiller grains with solubles in market turkey diets. Poult. Sci. 81(Supp. 1):92. Noll, S.L., V. Stangeland, G. Speers, C.M. Parsons and J. Brannon. 2003. Market tom turkey response to protein and threonine. Poult. Sci. 82 (Suppl. 1):73. Noll, S.L., J. Brannon and V. Stangeland. 2004. Market turkey performance and inclusion level of corn distillers dried grains with solubles. Poult. Sci. 83 (Suppl. 1):321 Noll, S.L., J. Brannon, J.L. Kalbfleisch and K.D. Roberson. 2005. Metabolizable energy value for corn distillers dried grains with solubles in turkey diets. Poult. Sci. 84 (Suppl. 1):12. Noll, S.L., and J. Brannon. 2006. Inclusion levels of corn distillers grains with solubles and poultry byproduct meal in market turkey diets. Poult. Sci. 85 (Suppl. 1):106-107. Roberson, K.D. 2003. Use of dried distillers grains with solubles in growing-finishing diets of turkey hens. Intl. J. Poult. Sci. 2(6):389-393. Waldroup P.W., J.A. Owen, B.E. Ramsey and D.L. Whelchel. 1981. The use of high levels of distillers dried grains plus solubles in broiler diets. Poult. Sci. 60:1479-1484.

99

USE OF ETHANOL DISTILLERS BYPRODUCTS IN LACTATING DAIRY COW DIETS David J. Schingoethe Dairy Science Department, Box 2104 South Dakota State University Brookings, SD 57007-0647 Phone: 605-688-5483 Fax: 605-688-6276 E-mail: david.schingoethe@sdstate.edu Summary Distillers grains with solubles (DGS) and corn gluten feed (CGF) are the major byproducts (coproducts) of ethanol production fed to cattle. This presentation discusses primarily results of feeding DGS to lactating cows, with some mention of feeding CGF and other ethanol coproducts. For both DGS and CGF, animal performance was usually similar when fed wet or dried products; however, some research results favored the wet products. Diets fed to dairy cattle can contain DGS or CGF as replacements for portions of both concentrates and forages, but usually replace concentrates. Distillers grains is a very good protein source (>30% CP) high in ruminally undegradable protein, and is very good energy source (NEL ~2.25 Mcal/kg of DM) for lactating cows. The modest fat concentration (10% of DM) and readily digestible fiber (40% NDF) contribute to the high energy in DGS. Distillers solubles are often blended with distillers grains to provide DGS, but the solubles can also be fed separately as "thin stillage" or as "condensed corn distillers solubles". Protein and energy values are similar for distillers grains with or without solubles but the phosphorus content is elevated when solubles are included. Extensive research has demonstrated that nutritionally balanced diets can be formulated that contain up to 20% of the diet dry matter as DGS. There is usually no advantage of feeding more than 20% DGS because the diet may contain excess protein and phosphorus, although production performance was very high even with more than 30% dried DGS in the diet. With more than 20% wet DGS in the diet, gut fill may decrease DM intake if other wet feeds are also in the diet. Milk composition is unchanged at all levels of DGS feeding, but fat content can be decreased if inadequate amounts of forage fiber are fed. Other coproducts from DGS processing such as condensed corn distillers solubles and corn germ can be well utilized in diets of lactating dairy cows. Corn gluten feed is a medium protein (24% CP) and medium energy (~1.73 Mcal NEL/kg of DM) feed that also contains an abundance of digestible fiber (35% NDF). While CGF can be fed at higher amounts than one usually feeds DGS, optimal production and feed efficiency of lactating cows occurred with 18 to 27% of ration DM as CGF. The fiber in DGS and CGF, which often replaces high starch feeds, does not eliminate acidosis but minimizes its problems. Innovations in processing technology will likely result in additional distillers coproducts for future use as livestock feeds. Nutrient Content of Ethanol Byproducts Nutrient content of the major ethanol coproducts is outlined in Table 1 with values listed usually for products from corn fermentation. These tabular values reflect primarily values reported in NRC (1996, 2001) as modified by more recently reported analytical information such as data from Spiehs et al. (2002) for new generation DGS and Birkelo et al. (2004) for the energy values of distillers grains. Such products tend to contain more protein, energy, and available phosphorus than distillers grains from older ethanol plants, which likely reflects increased fermentation efficiency in todays ethanol plants. Ethanol coproducts contain relatively high amounts of phosphorus, which can be a plus if additional phosphorus is needed in diets or a minus if excess phosphorus in manure needs to be disposed.

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

100

Table 1. Nutrient content of ethanol byproducts.1 DDGS2 Item Crude protein RUP5% of CP NEmaintenance, Mcal/kg NEgain, Mcal/kg NELactation, Mcal/kg Neutral detergent fiber (NDF) Acid detergent fiber (ADF) Ether extract Ash Calcium Phosphorus Magnesium Potassium Sodium Sulfur
1 2

Product Distillers CGF3 CGM4 solubles _____________________ (% of DM)__________________ 30.1 18.5 23.8 65.0 55.0 30.0 30.0 75.0 2.07 2.19 1.87 2.54 1.41 1.51 1.24 1.79 2.26 2.03 1.73 2.38 40.0 20.0 35.5 11.1 16.1 5.0 12.1 8.2 10.7 21.5 3.5 2.5 5.2 12.5 6.8 3.3 0.22 0.30 0.07 0.06 0.83 1.35 1.00 0.60 0.33 0.60 0.42 0.14 1.10 1.70 1.46 0.46 0.30 0.23 0.13 0.05 0.44 0.37 0.44 0.86

Most data are from NRC (1996, 2001), Spiehs et al. (2002), and Birkelo et al. (2004) DDGS = corn distillers grains 3 CGF = corn gluten feed 4 CGM = corn gluten feed 5 RUP = ruminally undegradable protein Virtually all distillers grains is marketed as distillers grains plus solubles, although this may change in the future as some processors fractionate distillers products into various components. The composition of corn distillers grains is essentially the same with or without solubles added, except for a lower phosphorus content (~0.4%) without solubles because the solubles are quite high in phosphorus (~1.35%). Therefore, most animal performance studies use data for distillers grains with or without solubles interchangeably. If a DGS product contains substantially more fat (e.g. >15%) and/or phosphorus (e.g. >1.0%) than the values listed in Table 1, it is very likely that more than normal amounts of distillers solubles were blended with the distillers grains, or that the processor had problems with separation of materials during the handling of solubles. Such variations also point out the importance of obtaining analytical data on the specific product being received from a supplier and the importance of suppliers providing uniform, standardized products. A complaint about some DGS suppliers is an inconsistent, variable product, a situation that is stimulating some other suppliers to offer consistent, premium quality DGS products that are sometimes even branded. Corn distillers grains is a good source of ruminally undegradable protein (RUP). The reported value of 55% of CP as RUP is probably an appropriate figure to use in most cases, although some variation in reported values exist. Most reported values range from 47% to 69% RUP (Firkins et al., 1984; Kleinschmit et al., 2007), with the higher quality products usually containing less than 64% RUP. Wet DGS usually has 5 to 8% lower concentrations of RUP than does dried DGS. Most of the readily degradable proteins in corn have been degraded during the fermentation process, thus the protein remaining in the corn DGS is going to be proportionately higher in RUP than in the original corn. However, if RUP values for DGS are quite high (e.g. >80% of CP), it may be advisable to check for heat damaged, undigestible protein. While some may wish to think that a golden yellow color is a good

101

indication of quality for DGS, research data from Belyea et al. (2004) indicated that color is sometimes (e.g. Powers et al., 1995) but often not (Kleinschmit et al., 2006b) an accurate indicator of protein quality. Distillers grains available in recent years contain more energy than older book values. Research by Birkelo et al. (2004) indicated that wet corn DGS contained approximately 2.25 Mcal/kg of NEL, 10 to 15% more energy than published in even the recent dairy NRC (2001) for dried DGS (DDGS). This likely reflects a higher energy value for newer generation distillers grains and does not necessarily reflect higher energy in wet than in dried DGS; that is a separate comparison that has not been made. Both DGS and CGF contain large amounts of NDF but low amounts of lignin. These readily digestible fiber sources can serve as partial replacements for forages as well as for concentrates in diets of dairy cattle. These nonforage fiber sources can supply energy needed for lactation or growth without the ruminal acid load caused by rapidly fermented starchy compounds (Ham et al., 1994). Such nonforage fiber sources of NDF can partially replace forages at times when forage supplies may be limited; however, because of the small particle size, DGS and wet CGF may lack sufficient effective fiber to prevent milk fat depression (Cyriac et al., 2005; Allan and Grant, 2000). Response of Lactating Cows Distillers Grains Milk production and composition data summarized in Table 2 are from the more than two-dozen research trials with 98 treatment comparisons reported between 1982 and early 2005 in which corn distillers grains, either wet or dried, were fed to lactating cows. This table is an abbreviated summary of the meta analysis conducted by Kalscheur (2005) of this extensive survey of virtually all of the modern research data available about feeding DGS to lactating cows. Amounts of DGS fed ranged from 4.2% of total diet DM (Broderick et al., 1990) to 41.6% of DM (Van Horn et al., 1985). Table 2. Dry matter intake (DMI), milk yield, milk fat, and protein content when fed diets containing wet or dried corn DGS.1 Inclusion level (% of DM) 0 4 10 10 20 20 30 > 30 SEM
a,b,c 1

DMI
_____________

Milk (kg/d)
_______________

Fat
_______________

Protein (%)
_______________

22.1b 23.7a 23.4ab 22.8ab 20.9c 0.8

33.0ab 33.4a 33.2ab 33.5a 32.2b 1.4

3.39 3.43 3.41 3.33 3.47 0.08

2.95a 2.96a 2.94a 2.97a 2.82b 0.06

Values within a column followed by a different superscript differ (P < 0.05). Adapted from Kalscheur (2005).

Production was the same as or higher when fed DGS than when fed control diets in virtually all experiments except possibly when fed very large amounts (i.e. 30% or more of diet DM) as wet DGS (Kalscheur, 2005). In experiments that compared DGS to soybean meal as the protein supplement, production was similar or higher, even when DDGS and soybean-based diets were formulated to be equal in RUP (Pamp et al., 2006). Florida research (Powers et al., 1995) indicated higher production when fed DDGS from whiskey or from fuel ethanol plants than when fed soybean meal. However, when they fed a DDGS product that was darker and possibly heat damaged, milk production was lower than when fed lighter, golden colored DDGS but still similar to production when fed soybean meal. But be cautioned because research data from Belyea et al. (2004) indicated that color is often not an accurate indicator of protein quality. When Kleinschmit et al. (2006b) used a standard, good quality DDGS to evaluate the response to two specially processed DDGS products intended to have even better quality, milk production

102

was higher for all three DDGS products than for the soybean meal-based control diet, with only small differences in response due to the improved DDGS quality. We are completing the second year of a trial in which cows were fed 15% of diet DM as wet DGS for the entire lactation, during the dry period, and into the second lactation. After the first year, there were no differences in production (31.7 and 33.6 kg/d for control and wet DGS), while fat percent (3.75 and 4.07), protein percent (3.29 and 3.41), and feed efficiency (1.30 and 1.57 kg FCM/kg DMI) were greater for cows fed wet DGS (Mpapho et al., 2006). Reproductive efficiency and cow health were similar for both dietary groups. The quality of protein in corn DGS is fairly good. As with most corn products, lysine is the first limiting amino acid in corn DGS for lactating cows, but corn DGS is a very good source of methionine. Therefore, sometimes (Nichols et al., 1998) but not always (Liu et al., 2000) milk production increased when fed supplemental ruminally protected lysine and methionine with DDGS, or when the DDGS was blended with other protein supplements that contained more lysine. While there may be differences in protein quality of various sources of DDGS present today (Kleinschmit et al., 2007), differences in yields of milk and milk protein might be slight, unless a product is greatly heat-damaged. Kleinschmit et al. (2006a) observed slightly greater production when 15% DDGS was fed in high alfalfa versus high corn silage diets, likely reflecting an improved amino acid status with the "blend" of alfalfa-DDGS proteins versus a diet containing predominantly corn-based proteins. Wet versus Dried DGS The response to wet or dried DGS is usually considered to be equal; however, very few trials actually compared wet versus dried DGS; most trials simply compared DGS to a control diet. When AlSuwaiegh et al. (2002) compared wet versus dried corn or sorghum DGS for lactating cows, they observed similar production for both wet and dried DGS but 6% more milk (P < 0.13) with corn versus sorghum DGS. Anderson et al. (2006) observed greater production when fed either wet or dried DGS than when fed the control diet, a tendency (P = 0.13) for greater production when fed wet DGS instead of dried DGS, and a tendency (P = 0.12) for greater production when fed 20% of the ration DM as DGS versus 10%, either wet or dried. The main considerations regarding the use of wet versus dried DGS are handling and costs. Dried products can be stored for extended periods of time, can be shipped greater distances more economically and conveniently than wet DGS, and can be easily blended with other dietary ingredients. Some possible problems with DDGS setting up when shipped extended distances in rail cars seems to be related to moisture and temperature conditions that some ethanol plants are addressing. Feeding wet DGS avoids the costs of drying the product, but there are other factors to consider with wet DGS that are not concerns when feeding dried DGS. Wet DGS will not remain fresh and palatable for extended periods of time; 5 to 7 days is the norm. Surface molds occasionally occur, thus there is usually some feed lost; a problem that wouldnt be a consideration with DDGS. The addition of preservatives may extend the shelf life of wet DGS by a few days (Spangler et al., 2005) but refereed journal publications that document such results are limited. We at SDSU (Kalscheur et al., 2002; 2003; 2004a,b) successfully stored wet DGS for more than six months in silo bags. The wet DGS was stored alone or blended with soyhulls (Kalscheur et al., 2002), with corn silage (Kalscheur et al., 2003), and with beet pulp (Kalscheur et al., 2004). Some field reports indicate successful preservation of wet DGS for more than a year in silo bags. Milk Composition When Fed DGS The composition of milk is usually not affected by feeding DGS unless routinely recommended ration formulation guidelines, such as feeding sufficient amounts of forage fiber, are not followed. Some field reports indicated milk fat depression when diets contained more than 10% of ration DM as wet DGS (Hutjens, 2004); however, those observations are not supported by research results. The meta analysis

103

(Kalscheur, 2005) showed no decreases in milk fat content when diets contained wet or dried DGS at any level, even as high as 40% of DM intake (see Table 2). In fact, the milk fat content was usually numerically highest for diets containing DGS. Incidentally, most of these studies were conducted during early to mid lactation, thus the data in Table 2 are typical for cows during these stages of lactation. The only time when milk fat content may have been lower with DGS was when diets contained less than 50% forage (Kalscheur, 2005). That result hints at why field observations of milk fat depression may have occurred. Because DGS contains an abundance of NDF, one may be tempted to decrease the amounts of forage fed when formulations indicate more than sufficient amounts of NDF. However, the small particle size of DGS means that its effective fiber is not as great as that of the forage fiber it replaced. Recent research at SDSU (Cyriac et al., 2005) support observations from the meta analysis. There was a linear decrease in milk fat concentration while milk production remained unchanged when cows were fed 0, 7, 14, and 21% of DM as DDGS in place of corn silage, even though dietary NDF content remained unchanged. The control diet contained 40% corn silage, 15% alfalfa hay, and 45% concentrate mix. Thus, the key to maintaining milk fat tests is to feed sufficient amounts of forage fiber. The fatty acid content of milk fat when cows are fed DGS is not expected to be affected greatly but has been evaluated in a few studies. Because the fat in DGS, especially corn DGS, is quite unsaturated with typically more than 60% linoleic acid, it is logical to expect a modest increase in concentrations of unsaturated fatty acids in the milk produced as observed by Schingoethe et al. (1999). Leonardi et al. (2005) and Anderson et al. (2006) also reported modest increases in the healthful fatty acid cis-9, trans-11 conjugated linoleic acid (CLA) and its precursor vaccenic acid (trans-11 C18:1). Milk protein content is seldom affected by feeding DGS unless protein is limiting in the diet. Then the lysine limitation in DGS may cause a slight decrease in milk protein content (Kleinschmit et al., 2006b). This effect may be more noticeable in diets that contain more than 30% DGS (Kalscheur, 2005), reflecting the high RUP and lysine limitation in DGS. How Much DGS can be Fed? We at SDSU and other researchers have demonstrated in several experiments that dairy producers can easily feed up to 20% of ration DM as distillers grains. With typical feed intakes of lactating cows, this is approximately 4.5 to 5.5 kg of dried DGS or 15 to 17 kg of wet DGS per cow daily. There are no palatability problems and one can usually formulate nutritionally balanced diets with up to that level of distillers grains in the diet using most combinations of forages and concentrates. For instance, with diets containing 25% of the dry matter as corn silage, 25% as alfalfa hay, and 50% concentrate mix, the DGS can replace most if not all of the protein supplement such as soybean meal and a significant amount of the corn that would normally be in the grain mix. This was illustrated in the experiment by Anderson et al. (2006) in which feeding 20% of the diet DM as wet or dried DGS replaced 25% of the corn and 87% of the soybean meal fed in the control diet. With diets that contain higher proportions of corn silage, even greater amounts of DDGS may be useable; however, the need for some other protein supplement, protein quality (e.g. lysine limitation), and phosphorus concentration may become factors to consider. With diets containing higher proportions of alfalfa, less than 20% DGS may be needed to supply the protein required in the diet. Thus, there are no strong advantages to feeding more than 20% distillers grains, but the possibility of feeding excess protein and/or phosphorus may occur. Grings et al. (1992) observed similar DM intake and milk production when cows were fed 31.6% of ration DM as DDGS. Schingoethe et al. (1999) fed slightly more than 30% of the ration DM as wet DGS with decreased DM intake but no decrease in milk production, likely reflecting the higher NEL content of the wet DGS diet. However, research by our group (Hippen et al., 2003; 2004) in which as much as 40% of ration DM was fed as DGS indicated possible problems when corn DGS provided more than 20 to 25% of the ration DM. Dry matter intake decreased with a corresponding decrease in milk production when wet DGS supplied more than 20% of the diet DM (Hippen et al., 2003). Gut fill may

104

have limited DM intake of these wet diets (40 to 46% DM) because total DM intake often decreases when the diet is less than 50% DM, especially when fermented feeds are fed (NRC, 2001). However, when dried DGS was fed (Hippen et al., 2004), DM intake and milk production still decreased when diets contained 27 to 40% DDGS. An interesting observation is that, in the meta analysis of 24 experiments (Kalscheur, 2005), the highest DM intakes and milk production occurred when diets contained 20 to 30% DGS although, as expected, DM intakes and production decreased with 30 to 40% wet DGS. Distillers Grains Blended with Other Feeds Several experiments were conducted at SDSU in which wet DGS was blended with other high fiber feeds. Such approaches may be helpful during times when forage supplies are limited or expensive. For instance, a 70:30 (DM basis) blend of wet DGS and soyhulls reduced the dustiness of soyhulls, reduced the seepage that is common with wet DGS, provided more desirable protein (21% CP) and P (0.6%) contents, and yet provided a high energy, high fiber feed (Kalscheur et al., 2002). We don't have lactation data on these forage blends but growth rates of heifers fed the blend were similar (1.22 and 1.27 kg/d) to gains when fed conventional diets (Kalscheur et al., 2004). Heifers fed a blend of wet DGS (69% of DM) and corn stalks (31%), had lower weight gains (1.04 kg/d) than when fed conventional diets (1.27 kg/d). Ensiling wet DGS alone or in combination with corn silage indicated that preservation of each could be enhanced by combining the feedstuffs with a 50:50 blend likely optimal (Kalscheur et al., 2003). Corn Distillers Solubles Distillers solubles are usually blended with the distillers grains before drying to produce DGS, but the solubles may be fed separately. Some include solubles in diets to decrease dustiness and decrease ingredient separation. DaCruz et al. (2005) fed 28% DM condensed corn distillers solubles (CCDS) at 0, 5, and 10% of total ration DM to lactating cows. Milk production (34.1, 35.5 and 35.8 kg/d for 0, 5, and 10% CCDS diets) increased when fed the CCDS, although milk fat (3.54, 3.33, and 3.43%) was slightly lower and milk protein (2.93, 2.97, 2.95%) was unaffected by diets. Recently, Sasikala-Appukuttan et al. (2006) fed as much as 20% of the total ration DM as CCDS (4% fat from the CCDS) with no apparent adverse affects on DM intake or milk composition. Milk yield tended to be higher for cows fed 10 and 20% CCDS than for cows fed the control diet. Thus, CCDS by itself can be a good feed for dairy cattle. However, we do not recommend feeding as much as 20% CCDS because diets including that much CCDS contained more than 0.5% phosphorus. Other Distillers Products One will see a growing list of new distillers products available as feeds for livestock in the future because processors continue to improve the efficiency of ethanol production and look for ways to fractionate byproducts resulting from the process. Abdelqader et al. (2006) recently completed an experiment feeding the germ that was removed from the corn grain prior to ethanol production. The germ (~21% fat) was fed to lactating cows at 0, 7, 14, and 21% of ration DM. Inclusion at 7 and 14% of DM increased milk and fat yields, however, feeding 21% corn germ decreased the concentration and yield of milk fat. Corn germ from wet milling operations may contain 45% or more fat, but feeding trials with that product are limited. A high protein (~45% CP) distillers grains product is coming available for evaluation. Ethanol producers are currently evaluating several approaches to remove fat from the coproducts for use in biodiesel. The results will be low fat DGS or other coproducts. Corn Gluten Feed Corn gluten feed, often fed as wet CGF, is a relatively high fiber, medium-energy, medium-crude protein product that can be fed to dairy cattle. The energy value of wet CGF was 92 to 100% of the energy value of shelled corn (Firkins et al., 1985; Ham et al., 1995); values were slightly lower for dry CGF. Lactating cows can consume quite large amounts of CGF with acceptable performance. Staples et

105

al. (1984) reported linear declines in DM intake and milk yield as amounts of wet CGF increased from at 0 to 40% of DM in 50% corn silage diets; however, dry matter content of the total diet may have been part of the problem as mentioned earlier regarding the feeding of wet DGS. Armentano and Dentine (1988) observed no reductions in DM intake and milk yield when diets contained as much as 7.9 kg/d (~36% of ration DM) as wet CGF. Wet CGF replaced only concentrates in most of the above studies. When wet CGF replaced up to 35% of ration DM as a mix of alfalfa hay, corn silage, and corn grain, milk production was greater than when fed the control diet (Van Baale et al., 2001). In experiments that included as much as 45% of ration DM as wet CGF, Schroeder (2003) concluded that 18.6% of dietary DM as wet CGF in place of portions of both forage and concentrate would maximize milk yield without negatively affecting milk composition or feed efficiency. Corn Gluten Meal Corn gluten meal (CGM) is a high protein (65% CP) high RUP (75% of CP) protein supplement; however, it is best to blend CGM with other protein supplements for optimal animal performance. Because of its high RUP level and lysine limitation, feeding CGM as the only protein supplement did not support the same amount of milk production as soybean meal-containing diets in a series of multiuniversity studies, even when the CGM diets were supplemented with ruminally protected lysine and methionine (Polan et al., 1991). A blend of several high quality proteins (blood meal, CGM, canola meal, and fish meal) supported milk production similar to production supported by soybean meal-containing diets (Piepenbrink et al., 1998). The Future? More new and improved distillers products will likely be available to the feed industry in the future. For instance, improvements in fermentation technology already provide DGS today that contains more protein and energy than DGS of previous years. It is also feasible to "fractionate" DGS into products that are higher in protein, other products that are higher in fat or in fiber, and products that are higher or lower in phosphorus (Rausch and Belyea, 2006). And some products from ethanol production may find their way into human foods and non-food products such as building products and biodiesel. For biodiesel, the fat may be extracted from the germ prior to ethanol fermentation, from the distillers solubles, or from the DGS. These comments are based on prior research experience with feeding whey, the coproduct from cheese manufacturing. At one time there was a choice between either liquid or dried "whole whey". Today, a large number of whey products varying from protein concentrates to lactose are available to the human food and animal feed industries. A similar situation could also occur with ethanol coproducts. References Abdelqader, M.M., A.R. Hippen, D.J. Schingoethe, K.F. Kalscheur, K. Karges, and M.L. Gibson. 2006. Corn germ from ethanol production as an energy supplement for lactating dairy cows. J. Dairy Sci. 89(Suppl. 1):156 (Abstr.) Allan, D.M., and R.J. Grant. 2000. Interactions between forage and wet corn gluten feed as sources of fiber in diets of lactating dairy cows. J. Dairy Sci. 83:322-331. Al-Suwaiegh, S., K.C. Fanning, R.J. Grant, C.T. Milton, and T.J. Klopfenstein. 2002. Utilization of distillers grains from the fermentation of sorghum or corn in diets for finishing beef and lactating dairy cattle. J. Anim. Sci. 80:1105-1111. Anderson, J.L., D.J. Schingoethe, K.F. Kalscheur, and A.R. Hippen. 2006. Evaluation of dried and wet distillers grains included at two concentrations in the diets of lactating dairy cows. J. Dairy Sci. 89:3133-3142. Armentano, L.E., M.R. Dentine. 1988. Wet corn gluten feed as a supplement for lactating dairy cattle and growing heifers. J. Dairy Sci. 71:990-995.

106

Belyea, R.L., K.D. Rausch, and M.E. Tumbleson. 2004. Composition of corn and distillers dried grains with solubles from dry grind ethanol processing. Biores. Technol. 94:293-298. Birkelo, C.P., M.J. Brouk, and D.J. Schingoethe. 2004. The energy content of wet corn distillers grains for lactating dairy cows. J. Dairy Sci. 87:1815-1819. Broderick, G.A., D.B. Ricker, and L.S. Driver. 1990. Expeller soybean meal and corn byproducts versus solvent soybean meal for lactating dairy cows fed alfalfa silage as the sole silage. J. Dairy Sci. 73:453462. Cyriac, J., M.M. Abdelqader, K.F. Kalscheur, A.R. Hippen, and D.J. Schingoethe. 2005. Effect of replacing forage fiber with non-forage fiber in lactating dairy cow diets. J. Dairy Sci. 88(Suppl. 1):252 (Abstr.) DaCruz, C. R., M. J. Brouk, and D. J. Schingoethe. 2005. Utilization of condensed corn distillers solubles in lactating dairy cow diets. J. Dairy Sci. 88:4000-4006. Firkins, J.L., L.L. Berger, and G.C. Fahey, Jr. 1985. Evaluation of wet and dry distillers grains and wet and dry corn gluten feeds for ruminants. J. Anim. Sci. 60:847-860. Firkins, J.L., L.L. Berger, G.C. Fahey, Jr., and N.R. Merchen. 1984. Ruminal nitrogen degradability and escape of wet and dry distillers grains and wet and dry corn gluten feed. J. Dairy Sci. 67:1936-1944. Grings, E.E., R.E. Roffler, and D.P. Deitelhoff. 1992. Response of dairy cows to additions of distillers dried grains with solubles in alfalfa-based diets. J. Dairy Sci. 75:1946-1953. Ham, G.A., R.A. Stock, T.J. Klopfenstein, and R.P. Huffman. 1995. Determining the net energy value of wet and dry corn gluten feed in beef growing and finishing diets. J. Anim. Sci. 73:353-359. Ham, G.A., R.A. Stock, T.J. Klopfenstein, E.M .Larson, D.H. Shain, and R.P. Huffman. 1994. Wet corn distillers byproducts compared with dried corn distillers grains with solubles as a source of protein and energy for ruminants. J. Anim. Sci. 72:3246-3257. Hippen, A.R., K.F. Kalscheur, D.J. Schingoethe, and A.D. Garcia. 2004. Increasing inclusion of dried corn distillers grains in dairy cow diets. J. Dairy Sci. 87: (Abstr.) Hippen, A.R., K.N. Linke, K.F. Kalscheur, D.J. Schingoethe, and A.D. Garcia. 2003. Increased concentration of wet corn distillers grains in dairy cow diets. J. Dairy Sci. 86(Suppl. 1):340 (Abstr.) Hutjens, M. F. 2004. Questions about wet distillers'. Hoard's Dairyman 149:261. Kalscheur, K.F. 2005. Impact of feeding distillers grains on milk fat, protein, and yield. Proc. Distillers Grains Technology Council, 10th Annual Symposium, Louisville, KY. Kalscheur, K.F., A.D. Garcia, A.R. Hippen, and D.J. Schingoethe. 2002. Ensiling wet corn distillers grains alone or in combination with soyhulls. J. Dairy Sci. 85(Suppl. 1):234 (Abstr.) Kalscheur, K.F., A.D. Garcia, A.R. Hippen, and D.J. Schingoethe. 2003. Fermentation characteristics of ensiling wet corn distillers grains in combination with corn silage. J. Dairy Sci. 86(Suppl. 1):211 (Abstr.) Kalscheur, K.F., A.D. Garcia, A.R. Hippen, and D.J. Schingoethe. 2004a. Growth of dairy heifers fed wet corn distillers grains ensiled with other feeds. J. Dairy Sci. 87:1964 (Abstr.) Kalscheur, K.F., A.D. Garcia, A.R. Hippen, and D.J. Schingoethe. 2004b. Fermentation characteristics of ensiled wet corn distillers grains in combination with wet beet pulp. J. Dairy Sci. 87(Suppl. 1):53 (Abstr.) Kleinschmit, D.H., J.L. Anderson, D.J. Schingoethe, K.F. Kalscheur, and A.R. Hippen. 2007. Ruminal and intestinal digestibility of distillers grains plus solubles varies by source. J. Dairy Sci. 90: (Submitted) Kleinschmit, D.H., D.J. Schingoethe, K.F. Kalscheur, and A.R. Hippen. 2006a. Evaluation of feeding dried distillers grains plus solubles (DDGS) with corn silage or alfalfa hay as the primary forage source. J. Dairy Sci. 89(Suppl. 1):109 (Abstr.) Kleinschmit, D.H., D.J. Schingoethe, K.F. Kalscheur, and A.R. Hippen. 2006b. Evaluation of various sources of corn distillers dried grains plus solubles for lactating dairy cattle. J. Dairy Sci. 89:47844794. Leonardi, C., S. Bertics, and L.E. Armentano. 2005. Effect of increasing oil from distillers grains or corn oil on lactation performance. J. Dairy Sci. 88:2820-2827.

107

Liu, C., D.J. Schingoethe, and G.A. Stegeman. 2000. Corn distillers grains versus a blend of protein supplements with or without ruminally protected amino acids for lactating cows. J. Dairy Sci. 83:2075-2084. Mpapho, G.S., A.R. Hippen, K.F. Kalscheur and D.J. Schingoethe. 2006. Lactational performance of dairy cows fed wet corn distillers grains for the entire lactation. J. Dairy Sci. 89:1811 (abstr.) National Research Council. 1996. Nutrient Requirements for Beef Cattle. 7th rev. ed. Natl. Acad. Sci., Washington, DC. National Research Council. 2001. Nutrient Requirements for Dairy Cattle. 7th rev. ed. Natl. Acad. Sci., Washington, DC. Nichols, J.R., D.J. Schingoethe, H.A. Maiga, M.J. Brouk, and M.S. Piepenbrink. 1998. Evaluation of corn distillers grains and ruminally protected lysine and methionine for lactating dairy cows. J. Dairy Sci. 81:482-491. Pamp, B.W., K.F. Kalscheur, A.R. Hippen, and D.J. Schingoethe. 2006. Evaluation of dried distillers grains versus soybean protein as a source of rumen-undegraded protein for lactating dairy cows. J. Dairy Sci. 89(Suppl. 1):403 (Abstr.) Piepenbrink, M.S., D.J. Schingoethe, M.J. Brouk, and G.A. Stegeman. 1998. Systems to evaluate the protein quality of diets fed to lactating cows. J. Dairy Sci. 81:1046-1061. Polan, C.E., K.A. Cummins, C.J. Sniffen, T.V. Muscato, J.L. Vincini, B.A. Crooker, J.H. Clark, D.G. Johnson, D.E. Otterby, B. Guillaume, L.D. Muller, G.A. Varga, R.A. Murray, and S.B. PierceSandner. 1991. Responses of dairy cows to supplemental rumen-protected forms of methionine and lysine. J. Dairy Sci. 74:2997-3013. Powers, W.J., H.H. Van Horn, B. Harris, Jr., and C.J Wilcox. 1995. Effects of variable sources of distillers dried grains plus solubles or milk yield and composition. J. Dairy Sci. 78:388-396. Rausch, K.D., and R.L. Belyea. 2006. The future of coproducts from corn processing. Applied Biochem. and Biotechnol. 128:47-85. Sasikala-Appukuttan, A.K., D.J. Schingoethe, A.R. Hippen, K.F. Kalscheur, K. Karges, and M.L. Gibson. 2006. The feeding value of corn distillers solubles for lactating dairy cows. J. Dairy Sci. 89(Suppl. 1) (Abstr.) Schingoethe. D.J., M.J. Brouk, and C.P. Birkelo. 1999. Milk production and composition from cows fed wet corn distillers grains. J. Dairy Sci. 82:574-580. Schroeder, J.W. 2003. Optimizing the level of wet corn gluten feed in the diet of lactating dairy cows. J. Dairy Sci. 86:844-851. Spangler, D., S. Gravert, G. Ayangbile, and D. Casper. 2005. Silo-King enhances the storage life and digestibility of wet distillers grains. J. Dairy Sci. 88:1922(Abstr.) Spiehs, M.J., M.H. Whitney, and G.C. Shurson. 2002. Nutrient data base for distillers dried grains with solubles produced from new generation ethanol plants in Minnesota and South Dakota. J. Anim. Sci. 80:2639-2645. Staples, C.R., C.L. Davis, G.C. McCoy, and J.H. Clark. 1984. Feeding value of wet corn gluten feed for lactating dairy cows. J. Dairy Sci. 67:1214-1220. Van Baale, M.J., J.E. Shirley, E.C. Titgemeyer, A.F. Park, M.J. Meyer, R.U. Lundquist, and R.T. Ethington. 2001. Evaluation of wet corn gluten feed in diets for lactating dairy cows. J. Dairy Sci. 84:2478-2485. Van Horn, H.H., O. Blanco, B. Harris, Jr., and D.K. Beede. 1985. Interaction of protein percent with caloric density and protein source for lactating cows. J. Dairy Sci. 68:1682-1695.

108

PROTEIN AND FERTILITY


James D. Ferguson, VMD, MS Professor, Clinical Nutrition Section of Animal Production Systems University of Pennsylvania, School of Veterinary Medicine, New Bolton Center 382 West Street Road Kennett Square, PA 19348 Phone: 610-925-6338 Fax: 610-925-8123 Email: ferguson@vet.upenn.edu Summary The effect of concentration of dietary crude protein (CP) on fertility of dairy cows has been reviewed by Ferguson and Chalupa, 1989 (J. Dairy Sci. 72:746-766.). In most studies, increasing dietary CP concentration has resulted in increased services/conception and days open, but responses have not been consistent across all studies. More and more evidence suggests that elevated protein per se is not the culprit causing reduced fertility, but elevations in rumen degradable protein, resulting in elevations in serum or plasma urea nitrogen, result in the reduction in fertility. More specifically, feeding rumen degradable protein above the amount needed for rumen synthesis of microbial protein is associated with reduced fertility. Effects appear to be due to changes in the uterine environment and oocyte development. Milk urea nitrogen can serve as a marker for potential negative effects on fertility. Milk urea nitrogen concentrations above 16 mg/dl may have negative associations with fertility and with increased environmental losses of N. The effect on fertility, although significant, is not a large magnitude in most circumstances, but environmental effects may be significant. Introduction Dietary crude protein is an important determinant of milk production. Under feeding CP is associated with reduced peak milk production. The partitioning of dietary CP into rumen degradable (RDP) and rumen undegradable (RUP) fractions has enabled a better understanding of protein utilization in the dairy cow. It has been recognized that feeding balances of rumen degradable (RDP) and rumen undegradable protein (RUP) consistent with requirements for rumen microbial synthesis and milk production can improve N utilization efficiency. In addition to improving milk production, providing sufficient balances of RDP and RUP may enhance fertility and reduce environmental losses of nitrogen (N). Protein and Fertility In general, increasing CP is dairy rations has been associated with reduced fertility, measured by increases in services/conception (reduction in conception rate (CR)) or days open (a measure of reproductive efficiency) (Ferguson and Chalupa, 1989). However, when studies were combined in a meta analysis, CP had no association with CR (Ferguson and Chalupa, 1989), but excess of RDP above that needed for rumen microbial synthesis was associated with reduction in CR (Figure 1).

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

109

Figure 1. Predicted CR (%) and milk urea nitrogen (MUN, mg/dl) as a function of increasing rumen ammonia balance. Range of rumen N balance calculated from protein and fertility studies in Ferguson and Chalupa, 1989. MUN predictions based on Roseler et al., 1993 and Hof et al. 1997. Rumen requirement for RDP is largely determined by fermentable carbohydrate. NRC estimates the requirement of RDP as 1.18 times the yield of microbial crude protein, which is assumed to be 130 g/kg total digestible nutrients (TDN) intake. This would equate to 24.5 g of degradable N per kg of TDN. Supplies of RDP providing more N per kg of TDN would increase rumen ammonia, plasma urea nitrogen (PUN) and milk urea nitrogen (MUN) concentrations (Figure 1). Milk urea nitrogen and PUN are highly correlated; relative differences in MUN and PUN concentrations will depend upon time of sampling PUN relative to feeding and sampling of MUN from composite or a.m.-p.m. milk samples. Milk urea nitrogen sampled from composite a.m.-p.m. milk samples tends to be a more stable estimate of mean urea nitrogen concentrations. Plasma and MUN will be used interchangeably in this paper. Ferguson et al. (1988) observed that fertility in a dairy herd was sensitive to elevated PUN. During periods when diets were offered with elevations in RDP, which increased PUN, CR declined in the herd. Cows with PUN greater than 20 mg/dl (3.3 mmol/l plasma urea, PU) had CR under 25% (Ferguson et al., 1988). The data suggested that PUN concentrations above 20.8 mg/dl were detrimental to fertility. Canfield et al. (1990) associated elevated PUN with reduced CR in an experiment with higher dietary RDP. Several studies have examined increasing PUN or MUN and CR in dairy cows (Ferguson et al., 1993; Butler et al., 1996; Rajala-Schultz et al., 2000; Melendez et al., 2000; Godden et al., 2001 and Hojman et al., 2004). A likelihood ratio test (LRT) for pregnancy was calculated for several of these studies and are presented in table 1. The LRT is calculated as the proportion of pregnant cows divided by the proportion of open cows within each urea category, as a proportion of the total cows. Pregnancy is more likely when the LRT is >1 and less likely when below one. In general, as urea concentration increases in plasma or in milk, fertility declines (Table 1). However, the decline in fertility is not uniform

110

across the studies, and the highest fertility group in Godden et al. (2001) was in the highest MUN category (Table 1). This table suggests that there is a general trend in reduction in fertility with increasing MUN, but MUN alone does not predict fertility. Multiple factors influence fertility. Other risk factors for fertility, not identified in these studies, may modify the association of increasing urea on fertility in dairy cows. Westwood et al. (2002) found that cows consuming increased RDP had lower fertility when associated with greater weight loss in the early postpartum period, suggesting energy balance may play a modifying role on nitrogen effects on fertility. Other factors may include body condition loss, metritis and earlier days of first insemination (Ferguson and Sklan, 2004). Table 1. Plasma or milk urea nitrogen (PUN, MUN, mg/dl) concentrations and likelihood ratio test (LRT) of pregnancy from Ferguson et al., 1993, Butler et al., 1996, Rajala-Schultz et al., 2000 and Godden et al., 2001.1

Ferguson et al., 1993 Butler et al., 1996 Rajala-Schultz et al., 20002 Godden et al.20013 PUN LRT PUN LRT MUN LRT MUN LRT --------------------------------------------------------------------------------------------------------------------<10 1.43 <16 2.65 <10 2.4 <9.8 1.12 10-14.9 1.01 16-18.9 1.61 10-14.9 1.4 9.8-12.59 1.11 15-19.9 0.90 19-21.9 0.81 15-19.9 1.2 12.6-15.39 0.96 20 - 24.9 0.92 22-24.9 0.80 20-24.9 1.0 15.4-18.2 0.85 >25 0.53 >25 0.73 >25 1.0 >18.2 1.17 --------------------------------------------------------------------------------------------------------------------Likelihood ratio test (LRT) calculated as the percent cows pregnant divided by percent cows open within PUN or MUN category. Ratios>1, pregnancy is more likely; ratios <1 pregnancy is less likely. 2 LRT calculated from risk ratios based on quartiles of MUN reported in Rajala-Schultz et al., 2000. Melendez et al. (2000) found negative associations of increasing MUN with fertility in summer versus winter months. Cows may adapt to high urea levels and maintain fertility (Westwood et al., 2002). Godden et al. (2001) found the relationship of fertility with urea was quadratic; fertility was higher in cows with low and high MUN concentrations (Table 1). Gustafsson et al. (1993) observed a similar relationship in Swedish herds. Increased MUN is correlated with increased urinary urea. Urinary urea breaks down rapidly to ammonia when mixed with feces. Ammonia volatilizes rapidly from barn floors and contributes to air particulate matter and acid rain. Therefore, reducing MUN has other benefits than reproduction. Together, the results suggest that fertility and environmental impact (and milk production) may be minimized when MUN concentrations are maintained between 9.0 to 16.0 mg/dl (1.7 and 2.7 mmol/l PU) on a herd basis. Individual cow concentrations may range from 4.0 to 22.0 mg/dl, but the majority of animals will cluster between 9.0 to 16.0 mg/dl. Thus, high production can be supported with adequate protein and minimal urea concentrations. Mechanisms Reducing Fertility Specific actions by which increasing urea concentrations associated with excess RDP reduce fertility have not been identified. Effects may be associated with alterations in the uterine environment which are detrimental to the early embryo or effects may detrimental to the oocyte, retarding development of the early blastocyst. Blanchard et al. (1990) observed embryo quality was reduced in cows consuming a 16.5% CP diet that contained 70% RDP compared with 62% RDP. The effect was not apparent in all cows, but particularly was seen in a higher proportion of cows in their 4th or greater parity. Approximately a third of cows consuming the higher RDP diet failed to yield any fertilized embryos. Larson et al., (1997) found that cows with higher MUN had more failed pregnancies which were
1

111

associated with regular inter-estrous intervals, based on sequential milk progesterone testing. These data suggest higher RDP and urea concentrations are associated with fertilization failure as a cause of repeat breeding and should result in regular inter-estrous intervals. However, Elrod et al. (1993a, 1993b) observed that reduced fertility with increasing serum urea nitrogen in heifers was associated with increased inter-estrous interval and reduction in uterine pH early in the luteal phase. Infertility was associated with increased embryonic loss. Elrods work suggested that loss of embryos occurred after maternal recognition of pregnancy, which extended the inter-estrous interval, resulting in reduced fertility. These results are in contrast to Blanchard et al. (1990) and Larson et al. (1997). Blanchard and Larsons studies were in lactating dairy cows whereas Elrods studies were in primiparous, nonlactating cows. Therefore mechanisms may be different. In addition, Blanchards study involved embryos collected from superovulated cows, seven days postinsemination whereas Larsons data was based on progesterone profiles postinsemination. Embryo loss prior to day 15 may have resulted in normal inter-estrous intervals in Larsons study. Sinclair et al. (2000) found higher dietary RDP increased serum ammonia and effected oocyte maturation and early blastocyst development. McEvoy et al. (1997) observed that plasma ammonia concentrations measured at or near insemination in sheep were negatively correlated (r = -0.27, p<.0002) with pregnancy. These studies suggest that increases in serum ammonia may play a role in reducing reproductive performance in cows fed high RDP diets by influencing oocyte quality and blastocyst maturation. DeWit et al. (2001) and Ocon and Hansen (2003) found that oocytes incubated in increasing concentrations of urea had reduced proportions of fertilized oocytes that developed to blastocysts. DeWit et al. (2001) found that increasing urea was associated with reduced fertilization and cleavage rate, but had no effect on embryos after fertilization. Ocon and Hansen (2003) reported that fewer oocytes developed to blastocysts due to decreased developmental competence. Urea reduced fertilization and cleavage rate of developing embryos. Armstrong et al. (2001) found increased urea associated with increased nutrient supply decreased oocyte quality. However, Lavan et al. (2004) observed that Holstein cows fed diets high in rapidly rumen degradable nitrogen experienced no negative effects on follicular development or embryo growth despite increases in serum urea and ammonia suggesting cows can adapt to short term increases in RDP. Few studies have examined the relationship between RUP and fertility. Westwood et al. (2000, 2002) concluded that increasing RUP in isonitrogenous diets improved feed intake, reduced serum nonesterified fatty acids postpartum, and improved reproductive performance particularly in cows of high genetic merit. Triplett et al. (1995) fed a basal diet to postpartum beef cows with three supplements of increasing RUP (low RUP, 38.1%; moderate RUP, 56.3%, and high RUP, 75.6%). Cows receiving the low RUP supplement had lower first service CR than cows receiving the moderate and high RUP supplement (29.2% versus 57.6% and 54.6%, respectively). Overall pregnancy proportion tended to be lower for the cows receiving the low RUP supplement than the moderate and high supplements (43.2%, 61.5% and 56.4%, respectively). It is difficult to separate the effects of increasing RUP on fertility from the simultaneous reduction in RDP which occurred in these studies. Conclusion Risk factors which modify N effects on fertility have not been clearly identified. Although, it seems fertility may be maintained at higher MUN concentrations, the general trend across the literature is a reduction in fertility. In addition, elevations in MUN are associated with increased urinary losses of N, a form of N which will be rapidly lost as ammonia to the environment. Nutritionists and veterinarians can monitor milk urea nitrogen (MUN) as a tool to assess efficiency of protein feeding. Mean MUN between 9.0 to 14 mg/dl is sufficient for adequate milk production and will ensure there are no negative effects on reproduction. Concentrations of MUN between 14 to 16 mg/dl should not significantly impair fertility but indicate some wastage of dietary N is occurring. MUN concentrations above 16 mg/dl not only may decrease fertility but also increase the risk of environmental pollution from ammonia volatilization.

112

References Armstrong, D.G., McEvoy, T.G., Baxter, G., Robinson, J.J., Hogg, C.O. 2001. Effect of dietary energy and protein on bovine follicular dynamics and embryo production in vitro: Associations with the ovarian insulin-like growth factor system. Biol. Reprod. 64:1624-1632. Blanchard, T., J. Ferguson, L. Love, T. Takeda, B. Henderson, J. Hasler, and W. Chalupa. 1990. Effect of dietary crude-protein type on fertilization and embryo quality in dairy cattle. Am. J. Vet. Res. 51:905908. Butler, W.R., J.J. Calaman, and S.W. Beam. 1996. Plasma and milk urea nitrogen in relation to pregnancy rate in lactating dairy cattle. J. An. Sci. 74:858-865. Canfield, R.W., C.J. Sniffen, and W.R. Butler. 1990. Effects of excess degradable protein on postpartum reproduction and energy balance in dairy cattle. J. Dairy Sci. 73:2342-2349. DeWit, A.A.C., M.L.F. Cesar, and T.A.M. Kruip. 2001. Effect of urea during in vitro maturation on nuclear maturation and embryo development of bovine cumulus-oocyte complexes. J. Dairy Sci. 84:1800-1804. Elrod, C.C. and W.R. Butler. 1993a. Reduction of fertility and alteration of uterine pH in heifers fed excess ruminally degradable protein. J. Anim. Sci. 71:694-701. Elrod, C.C., M. Van Amburgh, and W.R. Butler. 1993b. Alterations of pH in response to increased dietary protein in cattle are unique to the uterus. J. Anim. Sci. 71:702-706. Ferguson, J.D., T. Blanchard, D.T. Galligan, D.C. Hoshall, and W. Chalupa. 1988. Infertility in dairy cattle fed a high percentage of protein degradable in the rumen. J. Am. Vet. Assoc. 192:659-. Ferguson, J.D., and W. Chalupa. 1989. Impact of protein nutrition on reproduction in dairy cows. J. Dairy Sci. 72:746-766. Ferguson, J.D., D.T. Galligan, T. Blanchard, and M. Reeves. 1993. Serum urea nitrogen and conception rate: The usefulness of test information. J. Dairy. Sci. 76:3742-3746. Ferguson, J.D. and D. Sklan. 2005. Chapter 8. Effects of dietary phosphorus and nitrogen on cattle reproduction. Pages 233-253. In: Nitrogen and Phosphorus Nutrition in Cattle (A.N. Hristov and E. Pfeffer, eds.). CAB International. Godden, S.M., D.F. Kelton, K.D. Lissemore, J.S. Walton, K.E. Leslie, and J.H. Lusden. 2001. Milk urea testing as a tool to monitor reproductive performance in Ontario dairy herds. J. Dairy Sci. 84:13971406. Gustafsson, A.H. and J. Carlsson. 1993. Effects of silage quality, protein evaluation systems and milk urea content on milk yield and reproduction in dairy cows. Livest. Prod. Sci. 37:91-105. Hof, G., M.D. Vervoorn, P.J. Lenaers, and S. Tamminga. 1997. Milk urea nitrogen as a tool to monitor the protein nutrition of dairy cows. J. Dairy Sci. 80:3333-3340. Hojman, D., O. Kroll, G. Adin, M. Gips, B. Hanochi, and E. Ezra. 2004. Relationships between milk urea and production nutrition and fertility traits in Israeli dairy herds. J. Dairy Sci. 87:1001-1011. Larson, S.F., W.R. Butler, and W.B. Currie. 1997. Reduced fertility associated with low progesterone postbreeding and increased milk urea nitrogen in lactating cows. J. Dairy Sci. 80:1288-1295. Laven, R.A., P.M. Dawuda, R.J. Scaramuzzi, D.C. Wathes, H.J. Biggadike, and A.R. Peters. 2004. The effect of feeding diets high in quickly degradable nitrogen on follicular development and embryo growth in lactating Holstein dairy cows. Anim. Repro. Sci. 84:41-52. McEvoy, T.G., J.J. Robinson, R.P. Aitken, P.A. Findlay, and I.S. Robertson. 1997. Dietary excesses of urea influence the viability and metabolism of preimplantation sheep embryos and may affect fetal growth among survivors. Anim. Repro. Sci. 47:71-90. Melendez, P., A. Donovan and J. Hernandez. 2000. Milk urea nitrogen and infertility in Florida Holstein Cows. J. Dairy Sci. 83:459-463. Ocon, O.M. and P.J. Hansen. 2003. Disruption of bovine oocytes and preimplantation embryos by urea and acidic pH. J. Dairy Sci. 86:1194-1200. Rajala-Schultz, P.J., W.J.A. Saville, G.S. Frazer and T.E. Wittum. 2000. Association between milk urea nitrogen and fertility in Ohio dairy cows. J. Dairy Sci. 84:482-489.

113

Roseler, D.K., J.D. Ferguson, C.J. Sniffen, and J. Herrema. 1993. Dietary protein degradability effects on plasma and milk urea nitrogen and milk nonprotein nitrogen in Holstein cow. J. Dairy Sci. 76:525-534. Sinclair, K.D., M. Kuran, F.E. Gebbie, R. Webb and T.G. McEvoy. 2000. Nitrogen metabolism and fertility in cattle: II. Development of oocytes recovered from heifers offered diets differing in their rate of nitrogen release in the rumen. J. Ani. Sci. 78:2670-2680. Triplett, B.L., D.A. Neuendorff and R.D. Randel. 1995. Influence of undegraded intake protein supplementation on milk production, weight gain, and reproductive performance in postpartum Brahman cows. J. Ani. Sci. 73:3223-3229. Westwood, C.T., I.J. Lean, J.K. Garvin and P.C. Wynn. 2000. Effects of genetic merit and varying dietary protein degradability on lactating dairy cows. J. Dairy Sci. 83:2926-2940. Westwood, C.T., I.J. Lean and J.K. Garvin, J.K. 2002. Factors influencing fertility of Holstein dairy cows: A multivariate description. J. Dairy Sci. 85:3225-3237.

114

USING MINERAL AND VITAMIN SUPPLEMENTS TO ENHANCE REPRODUCTIVE PERFORMANCE OF DAIRY CATTLE James Spain Associate Professor of Animal Science University of Missouri Columbia Columbia, MO 65211 Phone: 573-882-0388 Email: spainj@missouri.edu Summary Reproductive function and fertility are essential to the efficient and profitable dairy enterprise. However, days from calving to conception and times bred per pregnancy on US dairy farms have continued a steady decline over the last several decades. The decline has been associated with the multiple factors involved in reproduction management. Increased confinement, larger herd sizes and less reliance on natural service have undoubtedly contributed to the decrease in reproduction efficiency. The level of milk production and the associated nutritional strain has also been implicated in having a negative impact on fertility and success of reproduction management systems. Most research has focused on the influence of energy and nitrogen balance on reproduction. These factors have been identified as especially important during the transition from late gestation through early lactation to the time of breeding. The greatest research emphasis has centered on energy balance and feeding management strategies to increase energy intake. More recent research has investigated the effects of micronutrients on fertility. One area of nutrition research that has shown promise is the supplementation of vitamins that improve energy metabolism related to lipid utilization. Rumen protected choline has been found to improve lipid utilization by the liver of transition dairy cows. The rumen protected choline was also shown to improve reproduction when fed to dairy cows during late gestation through breeding. Overall, vitamin and mineral status during the transition period from late gestation through the first 3 weeks of lactation have proven especially important. Improved selenium and Vitamin E balance has been shown to reduce risk of retained fetal membranes and therefore reduces the risk of associated infections and injury to the reproductive tract. Vitamin A and carotene have also shown to improve reproduction. Supplementation of diets of lactating dairy cows to improve energy metabolism while also maintaining mineral and vitamin homeostasis during transition and early lactation improves the fertility of the cow, resulting in improved reproduction.

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

115

USING FATS AND FATTY ACIDS TO ENHANCE REPRODUCTIVE PERFORMANCE


William W. Thatcher and Charles R. Staples Department of Animal Sciences University of Florida Gainesville, FL 32611 Phone: 352-392-5590 Fax: 352-392-5595 Email: thatcher@ufl.edu Summary It has been known for many years that early postpartum dairy cows usually produce more milk when fed a moderate amount of supplemental fat. There is growing evidence, as summarized in Table 2, that lactating dairy cows can benefit reproductively as well; although the responses are quite variable. Dietary fat effects on lactating dairy cows are not simply due to energy, but also due to specific fatty acids (e.g., linoleic acid, eicosapentaenoic acid, and conjugated linoleic acid) that interact as substrates for specific enzymes or interact with specific receptors (Peroxisome Proliferator-Activated Receptors) to regulate gene expression. Fat sources enriched in omega-6 or omega-3 fatty acids that deliver these fats to tissues beyond the rumen may be the most effective ones to feed but this can not be firmly concluded because other fats having very low amounts of these omega fatty acids have improved conception rates in single studies. The fats were fed at a minimum of 1.5% of the diet in studies in which conception rates were improved. Feeding less fat than this may be beneficial, but there is no supporting research behind it. Improved conception rates by fat-supplemented cows have been associated with an improved progesterone status of the cow by 1) increasing the size of the dominant follicle and corpus lutea on the ovaries and 2) by helping the corpus luteum survive and continue to produce progesterone during the early days of pregnancy. Early research results appear promising that some fat supplements may prove helpful to the health of the cow as well, but much more work is needed. If fed in moderate amounts, start feeding the fat when the cows enter the close-up group, especially if benefits to cow health and the ovaries are desired. Introduction Conception rates measured for cows managed under controlled experimental conditions as reported in scientific journals have decreased. Rates dropped from ~55% to ~45% (breeding at spontaneous estrus) to ~35% (timed AI) over a 50-year period (Lucy, 2001). Clearly reproductive performance of todays high producing dairy cow is sub-optimal with pregnancy rates estimated to be 20 % for the national herd (VanRaden et al., 2004) and 14.7% (de Vries and Risco, 2005) for the southeast region of the United States. The reasons for this infertility syndrome are multi-factorial, and seasonal periods of heat stress exacerbate the condition. Many reasons for reduced reproductive efficiency have been offered, including an increase in postpartum disease (ketosis, mastitis, retained fetal membranes, cystic ovaries, fatty liver, etc.), an increase in herd size resulting in increased management challenges, an increase in the proportion of milking heifers in the herd which cycle later, an increase in genetic inbreeding, and an increase in milk production (Lucy, 2001). Average milk production per lactation has increased by 57% from ~12,000 to ~19,000 pounds per cow in the last 25 years (Eastridge, 2006). However, amount of milk production has not been an accurate predictor of the chance for pregnancy. For example, higher-producing cows that ate very well cycled sooner after calving than lower-producing cows that ate poorly (Staples et al., 1990). Those poor eaters lost more body weight in the first 2 weeks postpartum and had not cycled by 9 weeks postpartum and fewer became pregnant. Extent of negative energy balance (energy output in milk plus
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

116

body maintenance minus energy intake from the diet) might be a more important factor influencing pregnancy. The lactating cow is more difficult to get pregnant than the nonlactating cow. The dollar value assigned to pregnancy for dairy cows varies because it is dependent upon several factors, such as how many days she has been milking, her lactation number, her milk yield, the replacement costs of a pregnant heifer, milk price, etc. A modeling program developed at the University of Florida was used to predict pregnancy value. Some key input values were a milk price of $14.09 per 100 pounds of milk, 305-d milk production of 23,144 lb for young animals and 25,994 lb for third lactation cows, and a replacement heifer cost of $1600 per head. The value of a new pregnancy at about 100 days in milk was calculated to be ~$200 for a milking heifer and ~$300 for a cow in her second lactation (de Vries, 2006). Even when a cow conceives, the pregnancy does not go to term about 50% of the time. If an average producing cow in the herd conceived at 61 days in milk, but was declared open 30 days later, the calculated loss ranged from $110 for heifers to $336 for cows in their third lactation (de Vries, 2006). Efforts to reduce this loss are certainly justified. The objective of this presentation is to evaluate the targeted influence of supplemental fat feeding as a nutraceutical approach to improve health and reproductive performance of lactating dairy cows. A nutraceutical is defined as a product isolated or purified from feeds that is demonstrated to have a physiological benefit or provide protection against chronic disease. Fats are more than just a source of energy in that specific classes of fatty acids exert direct regulatory effects on tissue function that impact milk production, immune status, and reproduction processes. The challenge is to understand these basic processes and to evaluate whether they have any benefit on reproductive efficiency. Fats Defined Many different types of supplemental fat have been fed to lactating cows. Some fat sources fed are listed in Table 1. Each fat source is composed of a different mix of individual fatty acids. Rendered fats include animal tallow and yellow grease (recycled restaurant grease) and are composed mainly of oleic acid (~43%). Granular fats are dry fats prepared commercially and are composed mainly of palmitic acid (36-50%). Examples include Energy Booster 100, EnerG-II, and Megalac-R. A variety of vegetable oils can be fed as free oil or in the seed form. The oil seeds contain from 18% oil (such as soybeans) to 40% oil (such as flaxseed). The selection of a vegetable oil will bring with it particular fatty acids. Canola oil is high in oleic acid. Cottonseed, safflower, sunflower, and soybean oils are high in linoleic acid. Flaxseed is high in linolenic acid. Linoleic acid and linolenic acid are essential fatty acids for the cow because neither her body nor her ruminal microorganisms can synthesize them. Fresh temperate grasses contain 1 to 3% fatty acids of which 55 to 65% is linolenic acid (Chilliard et al., 2001). Corn silage lipid contains much more linoleic acid (49%) than linolenic acid (4%) due to the presence of corn grain (Petit et al., 2004). Both linoleic and linolenic acid in forages can decrease during storage. As we have moved our dairy cows from pastures to barns and fed them stored forage, their intake of linolenic acid and possibly linoleic acid has likely decreased. The whole oil seed is frequently fed rather than the oil alone. Fish oil is unique in that it contains eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA), fatty acids found in fish tissue due to their consumption of marine plants. The short-hand notation for identifying fatty acids is to give the number of carbons and double bonds in the molecule. Fats that have double bonds are classified as unsaturated fats. For example, a designation of 18:2 indicates a fatty acid of 18 carbons long having 2 double bonds. The term omega refers to the location of the double bond in the carbon chain. An omega-6 fatty acid has its first double bond located between the 6th and 7th carbon counting from the methyl end of the chain. Likewise an omega-3 fatty acid has its first double bond located between the 3rd and 4th carbon counting from the methyl end of the carbon chain. Linoleic acid, abbreviated C18:2, is an omega-6 fatty acid. Linolenic acid, abbreviated C18:3, is an omega-3 fatty acid. Two additional omega-3 fatty acids are EPA (C20:5) and DHA (C22:6); but these are not considered essential for the cow because they can be synthesized

117

from the omega-3 fatty acid, linolenic acid. Nevertheless EPA and DHA can play important roles in supporting good animal performance. Table 1. Major fatty acid composition of select dietary fat sources. Fatty acid C14:0 C16:0 C16:1 C18:0 C18:1 Fat source Myristic Palmitic Palmit- Stearic Oleic oleic Tallow 3 25 3 18 43 Yellow grease 2 21 4 11 44 Energy Booster 100 1 3 40 1 41 10 Megalac; EnerG-II 1 1 50 <1 4 36 Megalac-R1 1 36 <1 4 26 Canola oil <1 4 <1 2 63 Cottonseed oil 1 23 1 3 18 Flaxseed oil <1 5 < 3 20 Extruded Linseed 7.6 5.2 20 Rapeseed oil <1 5 <1 2 54 Safflower oil <1 7 <1 2 12 Soybean oil <1 11 <1 4 23 Sunflower oil <1 7 <1 5 19 2 Menhaden fish oil 7 16 8 3 12 1 Commercial preparations considered partially inert in the rumen. 2 Also contains 14% C20:5 and 9% C22:6. Dietary Fats Are Modified in the Rumen by Bacteria The ruminal microbes will convert unsaturated fats to saturated fats by replacing the double bonds with single bonds between the carbons (called biohydrogenation). Some scientists have speculated that this act of biohydrogenation by bacteria is an attempt to protect themselves, as unsaturated fats can be toxic to bacteria, primarily the bacteria that digest fiber. The majority of the consumed unsaturated essential fatty acids, linoleic (C18:2) and linolenic (C18:3) acids, are converted by the bacteria to stearic acid (C18:0). During the process of biohydrogenation of unsaturated fats in the rumen, several intermediate forms of fatty acids, called trans fatty acids, also are formed. Some of the trans fatty acids, such as the trans-10, cis-12 conjugated linoleic acid (CLA) and the trans-10 C18:1, can influence the cows metabolism, including depressing milk fat synthesis. This intervention by ruminal bacteria to change essential fatty acids in the diet to other fatty acids has made the study of dietary fat effects on reproduction quite challenging. Evidence that Fatty Acids have Potential Intracellular Effects as Nutraceutical Effectors Dietary fat effects are not simply due to energy, but specific nutraceutical effects likely are being manifested whereby certain fatty acids interact as substrates for specific enzymes (e.g., PGHS-2) and also interact with peroxisome proliferator-activated receptors (PPARs) to regulate gene expression. The essential polyunsaturated fatty acids, linoleic (18:2n-6) and -linolenic (18:3n-3), undergo steps of chain elongation and desaturation forming differential n-6 products, such as dihomo-gamma linolenic (20:3n-6) and arachidonic (20:4n-6) acids, and n-3 products such as eicosapentaenoic acid (EPA; 20:5n-3). These specific long chain polyunsaturated fatty acids produce eicosanoid products of the prostaglandin series PGF1, PGF2 and PGF3, respectively as well as various thromboxanes, leukotrienes, lipoxins, hydroperoxy-eicosatetraenoic acids and hydroxyeicosatetraenoic acids (HETE) that regulate inflammation and immunity (Mattos et al., 1999). The peroxisome proliferator-activated receptors (PPARs) are a family

C18:2 Linoleic 3.8 14 2 8 29 19 54 16 14.5 22 78 54 68 1

C18:3 Linolenic <1 <1 <1 <1 3 9 1 55 51.3 11 <1 8 1 2

118

of nuclear receptors activated by specific fatty acids, eicosanoids and peroxisome proliferators. Previous studies have shown that they are involved in the regulation of genes affecting steroid and prostaglandin synthesis (MacLaren et al., 2006). We examined the effects of bST injection (500 mg; Posilac, Monsanto) and Ca salts of fatty acids enriched with fish oil on endocrine responses, ovarian-uterine function, expression of various genes in the uterus, and conceptus development on day 17 after estrus in cyclic and pregnant lactating dairy cows (Bilby et al., 2006a; Bilby et al., 2006b). Two diets were fed in which the oil of whole cottonseed (15% of dietary DM; control diet) was compared to oil prepared as a Ca salt containing fish oil (CaSFO) as one of the primary oils (1.9% of dietary DM; Virtus Nutrition). Formulated concentrations of ether extract (5.7%), nonstructural carbohydrate (36%), crude protein (18%), and net energy of lactation (1.7 Mcal/kg) were similar between diets (100% DM basis). Ingredients (corn silage, alfalfa hay, cottonseed hulls, ground corn, citrus pulp, soybean meal, expeller soybean meal, whole cottonseeds, calcium salt of fat, and mineral/vitamin premix) were fed as a totally mixed ration twice daily. All cows (n = 35) were fed the control diet for the first 10 days postpartum. Starting on day 11 of lactation, 8 of these cows were temporarily assigned to a transition diet containing 0.95% CaSFO to initially adjust cows to eating CaSFO. On day 18 postpartum, the CaSFO was increased to 1.9% of dietary DM on which they remained for the duration of the experiment. Ad libitum feed intakes were measured daily on a group basis. Daily feed intake by the cows fed CaSFO averaged 20.9 kg of DM/cow over the entire period of feeding. The calculated intake of EPA and DHA was 7.4 g/d for each fatty acid (i.e., 14.8 g/d total). Cows were milked thrice daily and individual milk weights recorded daily. All cows were presynchronized and started an Ovsynch protocol at approximately 7 days after a detected estrus. Cows assigned to bST treatment received an injection of bST at the Timed Artificial Insemination (TAI; i.e., cows were either TAI [pregnant] or not TAI [cyclic]) and 11 days later. The second injection at day 11 was done to insure that GH concentrations would be sustained until the time of slaughter on day 17. All cows that were designated for slaughter had to have ovulated and formed a CL when evaluated at day 7 following the second GnRH injection of the Ovsynch protocol. An important point is whether feeding a by-pass fat enriched in fish oil results in absorption of EPA and DHA that alters fatty acid concentrations among various tissues (Figure 1). The endometrium and liver had highest concentrations of C18:2n-6, C20:4n-6, C18:3n-3, EPA, and DHA compared to milk fat, mammary tissue, muscle, and both subcutaneous and internal fat tissues (Bilby et al., 2006c). An important observation was that the CaSFO diet reduced the concentrations of arachidonic acid (C20:4n-6) and preferentially increased the concentrations of EPA and DHA in the endometrium. Thus it is clear that EPA and DHA fatty acids of the CaSFO diet are being absorbed from the gastrointestinal tract and being preferentially taken up in the endometrial tissue.

Figure 1. Percents C18:2n-6, C20:4n-6, C18:3n-3, EPA and DHA of total fatty acids in milk fat, mammary, endometrium (endo), liver, muscle, subcutaneous (subQ) and internal fat tissues collected at day 17 of a synchronized estrous cycle.

119

Northern blot analyses indicated that PPAR and PPAR mRNAs were expressed in endometrium from cyclic and pregnant Holstein cows at day 17 following estrus. Western blot experiments confirmed the presence of PPAR and PPAR protein in bovine endometrium (MacLaren and Thatcher, in preparation). Supplementation with n-3 PUFArich fish oil was associated with reduced steady state concentrations of PPAR mRNA (Figure 2). In contrast, the CaSFO diet did not alter abundance of PPAR mRNA. Immunohistochemistry to localize PPAR indicated that the protein is Figure 2. Abundance of PPAR mRNA in endometrial expressed in the luminal epithelium, tissues at day 17 of a synchronized estrous cycle. in cyclic glandular epithelium, subepithelial stroma cows fed a control (Cyclic Con) diet, or cyclic cows fed a and, to a lesser extent, in the adluminal FO (Cyclic FO) diet, or pregnant cows fed a control diet stroma. The decrease in abundance of (Preg) with or without bST treatment at day 0 and 11 of PPAR mRNA associated with cows fed the synchronized estrous cycle. the CaSFO diet also was reflected with immunolocalization in that there was a reduction in moderate staining intensity of PPAR protein in the luminal epithelium of cows fed FO with or without bST treatment. This system undergoes an antiluteolytic pathway in the presence of a conceptus (i.e., decrease in ER- and oxytocin receptors) that leads to maintenance of the corpus luteum in pregnancy. In the present study, feeding CaSFO appeared to induce subtle antiluteolytic effects when examining immunohistochemistry spatial responses for the progesterone and estradiol- receptors and PGHS-2 proteins (Bilby et al., 2006b). Staining for progesterone receptors was evident in the superficial glandular epithelial cells of the cyclic cows at day 17 whereby CaSFO increased the moderate and heavy staining. Conversely, moderate to heavy staining intensity for the ER- receptor was reduced in the luminal epithelial cells of CaSFO-fed cows (Figure 3; Bilby et al., 2006b). This antiluteolytic pattern was associated with an attenuated effector response reflected by a decrease in heavy staining for PGHS-2 protein in the luminal epithelium of cyclic cows (Bilby et al., 2006b). Perhaps activation of the PPAR receptor normally attenuates expression of the progesterone receptor and enhances expression of the ER- receptor and PGHS-2 protein. Present results indicate that feeding by-pass fat enriched in fish oil (i.e., EPA and DHA) Figure 3. Percent staining (none, light, moderate and decreased PPAR receptors which would heavy) of estradiol- receptor protein in endometrial luminal epithelium collected on day 17 of a synchronized reverse these responses related to the estrous cycle in all (+/- bST) cyclic cows fed a control (All progesterone receptor, the ER- receptor Cyclic) or FO (All Cyclic FO) diets. and PGHS-2 protein. It is clear that EPA antagonizes the stimulatory effects of 120

arachidonic acid on PGF2 secretion (Mattos et al., 2003). As a consequence, a dietary manipulation of a nutraceutical has possibly altered the cellular and intracellular processes to sustain embryo development or pregnancy. The take home message is that feeding a by-pass fat enriched with fish oil appears to mediate specific effects in the uterus of cyclic cows that may benefit the processes of early pregnancy. The PPAR receptor that would bind the EPA ligand appears to be responsive to bST treatment in pregnant cows. Such results indicate that a by-pass fat diet can alter expression of a complement of genes in the uterus that impinge on uterine responses that support maintenance of pregnancy and development of the conceptus. An additional example of a potential fatty acid nutraceutical is the ability of the trans-10, cis-12 conjugated linoleic acid (CLA) to induce milk fat depression in lactating dairy cows. Utilizing a bovine mammary cell line (MAC-T), Peterson and coworkers (2004) indicate that trans-10, cis-12 CLA reduces lipid synthesis in the bovine mammary gland through inhibition of the proteolytic activation of sterol response element-binding protein (SREBP-1) that led to a subsequent reduction in transcriptional activation of lipogenic genes such as acetyl CoA carboxylase, fatty acid synthase, and stearoyl CoA desaturase. Consequently, supplemental diets enriched in polyunsaturated oils can cause a major reduction in milk fat content and this is due to the formation and absorption of trans-10, cis-12 CLA. Fat Supplementation and Reproductive Responses According to the scientific literature, a variety of fat supplements appear to benefit conception rates of lactating dairy cows (Table 2). The conception rates are sometimes reported for first insemination or for cumulated inseminations. Feedstuffs stimulatory to conception included calcium salts of palm oil distillate, tallow, Energy Booster (prilled tallow), MegaPro Gold (which is a calcium salt of palm oil plus rapeseed meal and whey permeate) fed to grazing cows, flaxseed (formaldehyde-treated or rolled), calcium salt of a mixture of soy oil and monounsaturated trans fatty acids), CLA, Megalac-R, and fish meal. The average improvement in conception rate was 18 percentage units. This is not to imply that the feeding of one of these feedstuffs to cows on your farm will increase herd conception rate by 18 percentage units. The margin of increase in conception rate due to fat feeding needed to be very great in these trials involving a few number of cows in order for the fat supplement to be detected as having a significant effect. If a fat supplement is to be beneficial on a dairy farm, it is more likely that the benefit will be less than 10 percentage units. In the 5 studies in which at least 250 cows participated in the study, the margin of increase due to fat supplementation averaged 8% with a range from -11 to 16 percentage units (Table 2). It is somewhat surprising that fat supplementation improved conception in experiments using between 30 and 132 cows. This may be partially due to the tighter management of cows used in experiments compared to that used on commercial farms. Normally about 300 cows per treatment are required to have a high chance of detecting a 10% increase in conception rate due to a treatment. Certainly other studies have been published in which fat-supplementation did not improve pregnancy rate. Staples et al. (1998) lists some of those studies, and some more recent studies are also presented in Table 2. In these studies, the fat-supplemented diet (last column in Table 2) was compared to a control diet that contained no supplemental fat in some studies, whereas in other studies the control diet contained another fat source. In head-to-head comparisons of fat sources, flaxseeds were more effective than rolled sunflower seeds (Ambrose et al., 2006a). Based on this, we might be tempted to conclude that the omega3 fatty acid (linolenic acid) found in flaxseed is more effective than the omega-6 fatty acid (linoleic acid) found in sunflower seeds. However, the linoleic acid in sunflower seeds is extensively biohydrogenated by the ruminal bacteria so that little is absorbed by the cow for reproductive use. This is based upon information that the linoleic acid in milk was not changed when cows were fed sunflower seeds (Petit et al., 2004). In two other head-to-head comparisons of fat supplements, cows fed calcium salts of palm oil distillate did not conceive as well as those fed either flaxseed (Petit et al., 2001) or those fed a calcium salt mixture of soybean oil and monounsaturated trans fatty acids (Juchem et al., 2004) (Table 2). Therefore fats containing more linolenic acid (flaxseed) or linoleic acid (soybean oil) may be more 121

effective than fats containing palmitic and oleic acid. However, as shown in Table 2, cows fed fats containing mainly palmitic and oleic acids such as tallow (Son et al., 1996) and Energy Booster (Frajblat and Butler, 2003) benefited cows reproductively. Possibly the trans fatty acids synthesized in the rumen by bacteria play a positive role for improving conception. Some of these trans fats can now be manufactured and fed to cows. The positive benefit to conception observed in Juchen et al., 2004 (34 vs. 26%) may have been due to the supplementation of the trans fat. Supplementing with trans fats (CLA) also appeared to benefit lactating cows in New York compared to those fed a calcium salt of palm oil distillate (81 vs. 44%; 42 vs. 27%) (Table 2). Increased synthesis of trans fats in the rumen is favored when cows are fed supplemental unsaturated fatty acids with high grain diets. Although the main nutrient in fish meal is protein and not fat, it is included here because there is growing evidence that the oils unique to fish may play a very important role in establishing pregnancy. The inclusion of fish meal in the diet (2.7 to 7.3% of dietary DM) also has improved either first service or overall pregnancy rate in four studies. In some of these studies, fish meal partially replaced soybean meal resulting in a reduction of an excessive intake of ruminally degradable protein. Therefore, the improved conception rates may have been due to the elimination of the negative effect of excessive intake of ruminally degradable protein on conception. However, in a field study in which the concentration of ruminally undegradable protein was kept constant between dietary treatments, cows fed fish meal had a better conception rate (Burke et al., 1997) suggesting that the positive response was due to something other than a reduction in intake of ruminally degradable protein. However, in the latter study conception rate was only improved in one of the two dairies, and it was the dairy with the greater level of milk production and an overall lower level of fertility. The unique omega-3 polyunsaturated fatty acids in fish (EPA and DHA) may have been responsible for the improvement in fertility, hence their inclusion in the current discussion. Oil seeds have not been well evaluated for their ability to improve conception. Those seeds that can deliver the key fatty acids past the rumen may be good candidates for the diet. Although the oil in many oil seeds contains more than 50% linoleic acid (Table 1), the delivery of linoleic acid past the rumen to the small intestine is not the same for all oil seeds. If we use an increase in the linoleic acid concentration of milk as an indicator that an oil seed can deliver linoleic acid to the tissues, then soybeans appear to be most effective and cottonseeds seem to be ineffective (Table 3). Sunflower seeds and safflower seeds also can increase the linoleic acid of milk fat, but not quite as effectively as that of soybeans. The processing of whole seeds also can influence their ability to deliver unsaturated fat past the rumen. Roasting of soybeans and rolling of sunflowers seemed to increase delivery of linoleic acid. Regarding linolenic acid, whole flaxseeds fed at about 10% of the diet can deliver some of its omega-3 fatty acid to the tissues. Grinding the flaxseed or heat-treated linseed (i.e., flaxseed) may deliver even more linolenic acid to the tissues (Table 3). In the United Kingdom, a process has been developed in which cracked linseeds or soybeans are processed with steam in order to create Maillard products, which help to protect the seeds unsaturated fatty acids from ruminal microbes (Robinson et al., 2002). Obviously, more research needs to be done to better identify the most effective fat sources, whether from seeds, oils, or calcium salts.

122

Table 2. Studies reporting conception rates (first service or cumulative services) of lactating dairy cows fed supplemental fatty acids. Unless otherwise indicated with a footnote, the control diet did not contain a fat supplement. Number of Control Fat treatment Reference Fat source cows in trial treatment Ferguson et al., 1990 2% Ca-Palm oil 253 43 591(P < 0.10) Sklan et al., 1991 Scott et al., 1995 Garcia-Bojalil et al., 1998 Son et al., 1996 2.6% Ca-Palm Oil 1 lb/d Ca-Palm oil 2.2% Ca-Palm oil 3% Tallow 99 443 43 68 81 30 121 309 356 129 397 42 32 30 132 80 62 300 62 93 52 44 582 503 324 37 393 35 263 27 443 273 52 44 68 32 45.5 82(P < 0.10) 98(P < 0.10) 86(P < 0.10) 62(P < 0.10) 86(P < 0.10) 87(P < 0.10) 481(P < 0.10) 261 39 54(P < 0.10) 341(P < 0.10) 581(P < 0.10) 81(P < 0.10) 42 72(P < 0.10) 64(P < 0.10) 89(P < 0.10) 41(P < 0.10) 63.5

Frajblat and Butler, 2003 1.7% Energy Booster Petit et al., 2001 Ambrose et al., 2006a Ambrose et al., 2006b Fuentes et al., 2007 McNamara et al., 2003 Juchem et al., 2004 Cullens et al., 2004 17% Flaxseed 9% Flaxseed 9% Flaxseed 1.7 kg/d Extruded Linseed 3.3 lb/d MegaPro Gold 1.5% Soy + Trans C18:1 2% Megalac-R

Castaneda-Gutierrez 0.3 lb/d Ca-CLA et al., 2005 Bernal-Santos et al.,2003 0.3 lb/d Ca-CLA Bruckental et al., 1989 Armstrong et al., 1990 Carroll et al., 1994 Burke et al., 1997 Average
1 2

7.3% Fish meal 1.8 lb/d Fish meal 3.5% Fish meal 2.8% Fish meal

First insemination. Control diet contained equal energy diet to fat-supplemented diet. Fat was fed prepartum only. 3 Control diet contained Ca salt of palm oil distillate. 4 Control diet contained rolled sunflower seeds.

123

Table 3. Effect of feeding various oilseeds on the essential fatty acid concentration of milk fat from dairy cows.1 Diet Reference Seed type Control + Oil Seed % of milk fatty acids C18:2 Dhiman et al., 1995 0% vs. 16% soybeans 3.2% 6.2%* Holter et al., 1992 Markus et al., 1996 Petit et al., 2004 Stegeman et al., 1992 Tice et al., 1994 Stegeman et al., 1992 C18:3 Petit et al., 2004 Gonthier et al., 2005 Fuentes et al., 2007 0% vs. 15% whole cottonseeds 0% vs. 7.1% whole sunflower seeds 0% or 9.6% whole sunflower seeds 0% or 10% rolled sunflower seeds 19.7% raw vs. roasted whole soybeans 0% or 10% rolled safflower seeds 0 vs.9.7% whole flaxseed 4.0% 2.3% 3.2% 2.2% 5.5% 2.2% 0.6% 4.2% 2.8%* 3.8% 3.3%* 6.7%* 3.1%* 1.1%*

0% vs. 12.5% ground flaxseed 0.4% 1.3%* 1% Ca Salt Palm Oil vs. 5.5% 0.5% 0.99%* Extruded Linseed * Values under the oilseed column having an asterisk were significantly different from the control values. An additional reproductive response is the specific evaluation of pregnancy losses following an early diagnosis of pregnancy. Feeding flaxseed to lactating dairy cows was reported to reduce pregnancy losses in two studies (0% for Flaxseed, 15.4% for Megalac, 8.0% for Micronized Soybeans [Petit and Twaginamungu, 2006]; 4.8% for Flaxseed and 11.4% for Sunflower seed; Ambrose et al., 2006a). However, a follow up study (Ambrose et al., 2006b) with a larger number of animals reported no differences in pregnancy losses between cows supplemented with Flaxseed (4/44 [9.1%] or without Flaxseed 9/63[14.3%]. Amount of Fat to Feed and When A frequently asked question is How much fat or of a specific fatty acid should be fed in order to improve reproduction? In the studies listed in Table 2, the fat sources were fed at a minimum of 1.5% of dietary dry matter. We know that feeding these amounts were effective. We do not know if feeding a smaller amount of fat would be effective as well. People are interested in feeding a smaller amount of fat for various reasons including keeping feed costs down and minimizing the potential negative effects of supplemental fats on the cows bacteria in her rumen. These negative effects can include reduced fiber digestion and reduced fat and protein concentrations in the milk. Generally speaking, fat supplementation at 1.5% of the diet is usually safe in terms of cow performance with the exception of fish oil. Feeding fish oil at more than 1% of dietary dry matter will usually reduce feed intake and/or milk fat and protein concentration. If the fat concentration of the base diet without a fat supplement is 3 to 4%, then increasing it to 4.5 to 5.5% by fat supplementation should not be a problem if the dietary fiber is sufficient and effective. Certainly diets containing higher fat ingredients like distillers grains, hominy, or whole cottonseeds need to be watched closely so that the total fat content stays below 6%. It is certainly possible that feeding supplemental fat at a lower rate such as 0.25 or 0.5 pounds per day could be effective. The key fatty acids (whether it is linoleic, linolenic, trans fatty acids, EPA, DHA, or something else) that do reach the small intestine of the cow are absorbed into the blood stream and

124

deposited into tissues including reproductive tissues. Some of these can accumulate over time. In a Florida study, the concentration of EPA increased in the liver fat from approximately 0.05 to 0.5 to 0.9% in liver samples collected at 2, 14, and 28 days in milk from cows fed linseed oil starting 5 weeks prepartum. A small but steady supply of these key fatty acids streaming to the tissues will allow the tissues to accumulate the fatty acids and have them ready at the proper time for reproductive purposes. Therefore, even a smaller fat feeding rate than the 1.5% as used by one of the published experiments in Table 2 could prove beneficial. Fat feeding must be initiated long enough before the fats are needed for restoring the reproductive tissues to a new fertile state. This would involve the involution of the uterus, the return of the ovaries to growing and ovulating new follicles, and the uterus to receiving and maintaining a new embryo successfully. As will be discussed later, cows fed selected fat sources have responded with larger (still of acceptable size) ovarian follicles. Since ovarian activity usually returns within the first 4 weeks of calving, initiating fat feeding prepartum would allow the absorbed fatty acids to influence early ovarian activity. Feeding supplemental fat for at least 21 days, preferably for 40 days, prior to the desired physiological response is our recommendation. We have begun supplementing cows in the close-up nonlactating period (3 to 5 weeks before the calculated due date). This allows the tissues to begin storing the key fatty acids prior to when they will be most needed. We conducted an experiment to test whether the initiation of fat supplementation (Megalac-R) (2% of dietary dry matter) should begin at 5 weeks prepartum, at calving, or at 28 days postcalving (Cullens et al., 2004). Cows fed fat starting in the prepartum period produced more milk and had fewer health problems in the first 10 days after calving than cows in the other groups. If some fat sources provide a benefit to the cows immune system, then the fat feeding should begin during the transition period. In summary, research studies that have documented the benefits of fat supplements for reproduction fed the fats at a rate of at least 1.5% of the diet dry matter. However, the reproduction responses are quite variable with some reports showing no significant benefit. Feeding fat supplements at a lower inclusion rate may prove beneficial in the field, but the research studies to support a lower feeding rate have not been done, as far as we know. How Might Fat Supplementation Help Improve Conception Rates? Improvements in reproductive performance through dietary fat supplementation may result for a variety of feasible reasons, either acting alone or in combination. The research support behind each of these is not equal, although each has a biological basis to be true. The main hypotheses are listed below as a group and then some are discussed individually in the following paragraphs. 1. The feeding of additional energy in the form of fat reduces the cows negative energy status so that she returns to estrus earlier after calving and therefore is more fertile at insemination. 2. Feeding additional essential fatty acids in the diet cures a fatty acid deficiency that has developed in the modern-day lactating dairy cow as she is managed today. 3. Cows fed fat develop larger ovarian follicles that develop into larger corpus lutea (CL) which produce more progesterone, a hormone necessary for coordinating nutrients for the developing embryo and for maintaining pregnancy until calving. 4. Specific individual fatty acids taken up by the uterus help inhibit the production or release of prostaglandin F2 (PGF2 ) by the uterus when the embryo is ~2.5 weeks old. This helps prevent the regression of the corpus luteum on the ovary so that progesterone continues to be produced and the newly formed embryo survives. Improving Energy Status? Those lactating dairy cows which experience a prolonged and intense negative energy state have a delayed resumption of estrous cycles after parturition which can increase the number of days open. If fat supplementation can help increase energy intake, then possibly the negative energy state can be lessened and estrous cycles start sooner and conception occur sooner. Adding a very energy dense nutrient such as

125

fat to the diet will usually increase the cows energy intake. However the energy status of the cow is usually not improved because of a slight to moderate depression in feed intake and/or an increase in milk production. Dairy cows fed tallow at 3% of dietary DM tended to have a greater pregnancy rate (62 vs. 44%; Table 2) despite having a more negative calculated mean net energy status from weeks 2 to 12 postpartum compared to cows not fed tallow. Likewise cows fed calcium salts of CLA (CastanedaGutierrez, 2005) or palm oil distillate (Garcia-Bojalil et al., 1998; Sklan et al., 1991) had better conception rates without an improvement in energy balance. Although there is evidence that the feeding of fat can improve the energy status of lactating dairy cows, an improvement in reproductive performance occurred in several instances apart from an improving energy status of the experimental animals. Therefore fat supplementation likely is improving reproductive performance by other alternative means as well. Healthier Ovarian Follicles? The size of the dominant follicle is often larger in lactating dairy cows receiving supplemental fat. On average, the size of the dominant follicle was 3.2 mm larger (a 23% increase) in fat-supplemented cows compared to control cows (Table 4). As shown in Table 4, a variety of dietary fat sources have had this effect on cow ovaries. Yet are certain fats more effective? Some studies did compare fat sources head-to-head. In two studies, it was the feeding of fats enriched in omega-6 (linoleic acid) or omega-3 fatty acids (linolenic or EPA and DHA) (Staples et al., 2000; Bilby et al., 2006d) that stimulated larger dominant follicles compared to fats enriched in oleic acid. Thus the polyunsaturated fats were most effective in increasing follicle size. Thus, cows fed fats enriched with the essential fatty acids are likely to have more progesterone being synthesized due to a formation of a larger ovarian corpus luteum that is derived from a larger ovulatory follicle. Potential increases in progesterone secretion and plasma concentrations need to be considered with regard to the higher rate of metabolism and clearance of progesterone in lactating dairy cows. Our recent study of feeding calcium salts of Fish Oil, beginning early postpartum, had no positive or negative effect on progesterone concentrations during a programmed estrous cycle (Bilby et al., 2000a). Furthermore, Moussavi et al. (2007) failed to detect any affect of feeding either fishmeal or calcium salts of Fish Oil on plasma progesterone concentrations during a programmed estrous cycle until day 15. Antiluteolytic Effect of Reducing Prostaglandin Secretion? If PGF2 is released by the uterus, the corpus luteum will regress, progesterone synthesis will decline, the embryo will die for lack of support, and the cow will start a new estrous cycle. About 50% of embryos die (~40% during the first 28 days after AI and ~14% between 28 and 45 days after AI). Embryonic loss is a significant problem in the dairy industry. Omega-3 fatty acids stored in the uterus from the diet may aid the process of embryo development and survival by helping to reduce the synthesis of prostaglandin F2. Quite clearly EPA, DHA, and -linolenic fatty acids can inhibit PGF2 release by bovine endometrial cells in vitro. Can omega-6 fatty acids have a similar beneficial effect? Not likely, because omega-6 fatty acids are used to synthesize prostaglandin F2; although CLA can suppress PGF2 release in vitro. Lactating dairy cows fed soybeans or sunflower seeds (both good sources of linoleic acid, the omega-6 fatty acid) had increased concentrations of prostaglandin F2 in their blood when the uterus was artificially stimulated with an oxytocin injection. Cows that are fed omega-3 fatty acids partially replace the omega-6 fatty acids stored in the uterus so that there is less omega-6 inventory for the cow to draw from for synthesis of prostaglandin F2. In demonstration of this effect, a series of studies demonstrated: feeding fish meal reduced the PGFM response to an estradiol-oxytocin challenge (Mattos et al., 2002); cows fed sunflower seeds (enriched in linoleic acid) had higher PGFM response compared to cows fed linseed, Megalac or no fat supplement (Petit et al., 2004); cows fed linoleic or -linolenic fatty acids did not differ in their PGFM response to oxytocin on days 15 and 16, but response to linoleic was greater on day 17 (Robinson et al., 2002). Moussavi et al., (2007) failed to detect a difference in PGFM response to oxytocin at day 15 of a programmed estrous cycle in the early postpartum period (i.e., 126

d 49 DIM) when either Fish meal or calcium salts of Fish Oil were fed previously. Variability in these results may be associated with timing of the induction response relative to normal time of luteolysis. Alternative models need to be examined to conclusively demonstrate the antiluteolytic effect of omega-3 fatty acids on prostaglandin secretion. Table 4. Diameter of the dominant ovarian follicle of lactating dairy cows fed fat supplements was greater than that of cows fed the control diet (P < 0.10). Experimental diets Reference Lucy et al., 1991 (see Staples et al., 1998) Lucy et al., 1993 (see Staples et al., 1998) Oldick et al., 1997 (see Staples et al., 1998) Beam and Butler, 1997 (see Staples et al., 1998) Staples et al., 2000 Robinson et al., 2002 Bilby et al., 2006 Ambrose et al., 2006 Average References Ambrose, D.J., C.T. Estill, M.G. Colazo, J.P. Kastelic and R. Corbett. 2006b. Conception rates and pregnancy losses in dairy cows fed a diet supplemented with rolled flaxseed. Proc 7th International Ruminant Reproduction Symposium, Wellington, New Zealand. Abstract 50. Ambrose, D.J., J.P. Kastelic, R. Corbett, P.A. Pitney, H.V. Petit, J.A. Small and P. Zalkovic. 2006a. Lower pregnancy losses in lactating dairy cows fed a diet enriched in -linolenic acid. J. Dairy Sci. 89:3066-3074. Armstrong, J.D., E.A. Goodall, F.J. Gordon, D.A. Rice and W.J. McCaughey. 1990. The effects of levels of concentrate offered and inclusion of maize gluten or fish meal in the concentrate on reproductive performance and blood parameter of dairy cows. Animal Production 50:1-10. Bernal-Santos, G., J.W. Perfield II, D. M. Barbano, D.E. Bauman and T.R. Overton. 2003. Production responses of dairy cows to dietary supplementation with conjugated linoleic acid (CLA) during the transition period and early lactation. J. Dairy Sci. 86:3218-3228. Bilby, T.R., A. Guzeloglu, L.A. MacLaren, C.R. Staples and W.W. Thatcher. 2006a. Pregnancy, bST and omega-3 fatty acids in lactating dairy cows: II. Gene expression related to maintenance of pregnancy. J. Dairy Sci. 89:3375-3385. Bilby, T.R., A. Sozzi, M.M. Lopez, F. Silvestre, A.D. Ealy, C.R. Staples and W.W. Thatcher. 2006b. Pregnancy, bST and omega-3 fatty acids in lactating dairy cows: I. ovarian, conceptus and growth hormone IGF system response. J. Dairy Sci. 89:3360-3374. Fat source Control Fat

-------- mm ------Ca salt of palm oil Ca salt of palm oil Yellow grease Tallow Yellow grease Soybean oil, fish oil Protected soybeans Megalac-R or Flaxseed oil Rolled flaxseeds 12.4 16.0 16.9 11.0 14.3 13.3 15.0 14.1 14.1 18.2 18.6 20.9 13.5 17.1 16.9 16.5 16.9 17.3

127

Bilby, T.R., J. Block, J.O. Filho, F.T. Silvestre, B.C. Amaral, P.J. Hansen, C.R. Staples and W.W. Thatcher. 2006d. Effects of dietary unsaturated fatty acids on oocyte quality and follicular development in lactating dairy cows in summer. J. Dairy Sci. 89:3891-3903. Bilby, T.R., T. Jenkins, C.R. Staples and W.W. Thatcher. 2006c. Pregnancy, bST and omega-3 fatty acids in lactating dairy cows: III. Fatty acid distribution. J. Dairy Sci. 89:3386-3399. Bruckental, I., D. Dori, M. Kaim, H. Lehrer and Y. Folman. 1989. Effects of source and level of protein on milk yield and reproductive performance of high-producing primiparous and multiparous dairy cows. Animal Production. 48:319-329. Burke, J.M., C.R. Staples, C.A. Risco, R.L. De La Sota and W.W. Thatcher. 1996. Effect of ruminant grade menhaden fish meal on reproductive and productive performance of lactating dairy cows. J. Dairy Sci. 80:3386-3398. Carroll, D.J., F.R. Hossain and M.R. Keller. 1994. Effect of supplemental fish meal on the lactation and reproductive performance of dairy cows. J. Dairy Sci. 77:3058-3072. Castaneda-Gutierrez, E., T.R. Overton, W.R. Butler and D.E. Bauman. 2005. Dietary supplements of two doses of calcium salts of conjugated linoleic acid during the transition period and early lactation. J. Dairy Sci. 88:1078-1089. Chilliard, Y., A. Ferlay and M. Doreau. 2001. Effect of different types of forages, animal fat or marine oil in cows diet on milk fat secretion and composition, especially conjugated linoleic acid (CLA) and polyunsaturated fatty acids. Livestock Production Science 70:31-48. Cullens, F.M., C.R Staples, T.R. Bilby, F. Silvestre, J. Bartolome, A. Sozzi, L. Badinga, W.W. Thatcher and J.D. Arthington. 2004. Effect of timing of initiation of fat supplementation on milk production, plasma hormones and metabolites, and conception rates of Holstein cows in summer. J. Dairy Sci. 86(Suppl. 1):308. De Vries, A., and C.A. Risco. 2005. Trends and Seasonality of Reproductive Performance in Florida and Georgia Dairy Herds from 1976 to 2002. J. Dairy Sci. 88:3155-3165. De Vries, A. 2006. Economic value of pregnancy in dairy cattle. J. Dairy Sci. 89:3876-3885. Dhiman T.R., K.V. Zanten and L.D. Satter. 1995. Effect of dietary fat source on fatty acid composition of cows milk. Journal of the Science of Food and Agriculture 69:101-107. Eastridge, M.L. 2006. Major Advances in Applied Dairy Cattle Nutrition. J. Dairy Sci. 89:1311-1323. Ferguson, J.D., D. Sklan, W.V. Chalupa and D.S. Kronfeld. 1990. Effect of hard fats on in vitro and in vivo rumen fermentation, milk production, and reproduction in dairy cows. J. Dairy Sci. 73:2864-2879. Frajblat, M., and W.R. Butler. 2003. Effect of dietary fat prepartum on first ovulation and reproductive performance in lactating dairy cows. J. Dairy Sci. 86(Suppl. 1):119. Fuentes, M.C., S. Calsamiglia, C. Sanchez, A. Gonzalez, J.R. Newbold, J.E.P. Santos, L.M. RodriguezAlcala and J. Fontecha. Effect of extruded linseed on productive and reproductive performance of lactating dairy cows. Livestock (In Press, via J.E.P. Santos). Garcia-Bojalil, C.M., C.R. Staples, C.A. Risco, J.D. Savio and W.W. Thatcher. 1998. Protein degradability and calcium salts of long-chain fatty acids in the diets of lactating dairy cows: Reproductive responses. J. Dairy Sci. 81:1385-1395. Gonthier, C., A.F. Mustafa, D.R. Ouellet, P.Y. Chouinard, R. Berthiaume and H.V. Petit. 2005. Feeding micronized and extruded flaxseed to dairy cows: effects on blood parameters and milk fatty acid composition. J. Dairy Sci. 88:748-56. Juchem, S.O., R.L.A. Cerri, R. Bruno, K.N. Galvao, E.W. Lemos, M. Villasenor, A.C. Coscioni, H.M. Rutgliano, W.W. Thatcher, D. Luchini and J.E.P. Santos. 2004. Effect of feeding Ca salts of palm oil (PO) or a blend of linoleic and monoenoic trans fatty acids (LTFA) on uterine involution and reproductive performance in Holstein cows. J. Dairy Sci. 87(Suppl. 1):310. Lucy, M.C. 2001. Reproductive loss in high-producing dairy cattle: Where will it end? J. Dairy Sci. 84:1277-1293. MacLaren, L.A., A. Guzeloglu, A.F. Michel, and W.W. Thatcher. 2006. Peroxisome proliferator-activated receptor (PPAR) expression in cultured bovine endometrial cells and response to omega-3 fatty acid,

128

growth hormone and agonist stimulation in relation to series 2 prostaglandin production. Dom. Anim. Endo. 30:155-169. Markus, S.B, K.M. Wittenberg, J.R. Ingalls and M. Undi. 1995. Production responses by early lactation cows to whole sunflower seed or tallow supplementation of a diet based on barley. J. Dairy Sci. 79:1817-1825. Mattos, R, C.R. Staples and W.W. Thatcher. 1999. Effects of dietary fatty acids on reproduction in ruminants. Rev. Reprod. 5:38-45. Mattos, R., A.Guzeloglu, L. Badinga, C.R. Staples and W.W. Thatcher. 2003. Polyunsaturated fatty acids and bovine interferon- modify phorbol ester-induced secretion of prostaglandinF2 and expression of prostaglandin endoperoxide synthase-2 and phospholipase-A2 in bovine endometrial cells. Biol. Reprod 69:780-787. McNamara, S., T. Butler, D.P. Ryan, J.F. Mee, P. Dillon, F.P. OMara, S.T. Butler, D. Anglesey, M. Rath and J.J. Murphy. 2003. Effect of offering rumen-protected fat supplements on fertility and performance in spring-calving Holstein-Friesian cows. Anim. Reprod. Sciences 79:45-56. Moussavi, A.R., R.O. Gilbert, T.R. Overton, D.E. Bauman and W.R. Butler. 2007. Effects of feeding fish meal and n-3 fatty acids on ovarian and uterine responses in early lactating dairy cows. J. Dairy Sci. 90:145-154. Peterson, D.G., E.A. Matitashvili and D.E. Bauman. 2004. The Inhibitory Effect of trans-10, cis-12 CLA on Lipid Synthesis in Bovine Mammary Epithelial Cells Involves Reduced Proteolytic Activation of the Transcription Factor SREBP-1. J. Nutr. 134:2523-2527. Petit, H.V., and H. Twagiramungu. 2006. Conception rate and reproductive function of dairy cows fed different fat sources. Theriogenology 66:1316-1324. Petit, H.V., R.J. Dewhurst, J.G, Proulx, M. Khalid, W. Haresign and H. Twagiramungu. 2001. Milk production, milk composition, and reproductive function of dairy cows fed different fats. Canadian J. Dairy Sci. 81:263-271. Petit, H.V., C. Germiquet and D. Lebel. 2004. Effect of feeding whole, unprocessed sunflower seeds and flaxseed on milk production, milk composition, and prostaglandin secretion in dairy cows. J. Dairy Sci. 87:3889-3898. Robinson, R.S., P.G.A Pushpakumara, Z. Cheng, A.R. Peters, D.E.E. Abayasekara and D.C. Wathes. 2002. Effects of dietary polyunsaturated fatty acids on ovarian and uterine function in lactating diary cows. Reproduction 124:119-131. Scott, T.A., R.D. Shaver, L. Zepeda, B. Yandell and T.R. Smith. 1995. Effects of rumen-inert fat on lactation, reproduction, and health of high producing Holstein herds. J. Dairy Sci. 78:2435-2451. Sklan, D., U. Moallem and Y. Folman. 1991. Effect of feeding calcium soaps of fatty acids on production and reproductive responses in high producing lactating cows. J. Dairy Sci. 74:510-517. Son, J., R.J. Grant and L.L. Larson. 1996. Effects of tallow and escape protein on lactational and reproductive performance of dairy cows. J. Dairy Sci. 79:822-830. Staples, C.R., J.M. Burke and W.W. Thatcher. 1998. Influence of supplemental fats on reproductive tissues and performance of lactating cows. J. Dairy Sci. 81:856-871. Staples, C.R., W.W. Thatcher and J.H. Clark. 1990. Relationship between ovarian activity and energy status during the early postpartum period of high producing dairy cows. J. Dairy Sci. 73:938-947. Staples, C.R., M.C. Wiltbank, R.R. Grummer, J. Guenther, R. Sartori, F.J. Diaz, S. Bertics, R. Mattos and W.W. Thatcher. 2000. Effect of long chain fatty acids on lactation performance and reproductive tissues of Holstein cows. J. Dairy Sci. 83(Suppl. 1):278. Stegeman, G.A., R.J. Baer, D.J. Schingoethe and D.P. Casper. 1992. Composition and flavor of milk and butter from cows fed unsaturated dietary fat and receiving bovine somatotropin. J. Dairy Sci. 75:962970. Tice, E.M., M.L. Eastridge and J.L. Firkins. 1994. Raw soybeans and roasted soybeans of different particle sizes. 2. Fatty acid utilization by lactating cows. J. Dairy Sci. 77:166-180. VanRaden, P.M., A.H. Sanders, M.E Tooker, R.H. Miller, H.D. Norman, M.T. Kuhn and G.R.Wiggans. 2004. Development of a national genetic evaluation for cow fertility. J. Dairy Sci. 87:2285-2292. 129

NUTRITIONAL MANAGEMENT OF THE ORGANIC DAIRY Robert C. Fry, DVM Atlantic Dairy Management Services Kennedyville, MD 21645 Phone: 410-652-5538 Fax: 410-648-5353 Email: rcfry@baybroadband.net Summary Management of cows in certified organic dairies requires an understanding of organic protocols and regulations for nutritionists, veterinarians and producers. If a dairyman chooses to manage his or her cows under organic guidelines, specific principles of management that are unique to organic production must be coupled with best-known traditional management practices used in non-organic herds. Introduction Professional publications, consumer magazines, newspapers, and World Wide Web sights discussing the merits of organic food for human consumption are abundant. Consumer access to information about the details surrounding the food they eat combined with ever more frequent news stories about unsafe food has fueled an unprecedented growth in the organic food industry. Since 1997, organic food sales have increased each year by 17-21 percent while during the same period total food sales in the US have only climbed 2-4 percent. In 2001 the organic movement in the United States was a $7.7 billion business and in 2006, $40 billion according to research by Organic Monitor. Dairy products are no exception to this with a 17% growth in sales of organic dairy products expected through the year 2008 (Martin, 2005). Continuing consumer demand for organic food has milk processors clamoring for a supply of certified organic milk. Discussing the merits of this booming business is not in the scope of this paper. Instead the focus will be on understanding organic protocols so nutritionists and veterinarians can service the needs of clients who choose to manage their dairy for production of certified organic milk. Discussion Understanding organic dairy production requires some review of the organic food regulations in our country. In 1990 the US Congress passed The Organic Foods Production Act as part of the Farm Bill. This act authorizes the Secretary of Agriculture to appoint a 15 member advisory committee called the National Organic Standards Board (NOSB). The board serves as an advisor to the Secretary of Agriculture regarding the implementation of the United States Department of Agricultures (USDA) National Organic Program (NOP) and assisting in developing standards for materials used in organic production as listed by the NOP. The NOP regulations are a 544-page document published in the Federal Register under the direction of the Agricultural Marketing Service (AMS), an arm of the USDA. This national program facilitates domestic and international marketing of fresh and processed food that is organically produced, assuring consumers of consistent and uniform production standards. Additionally the NOP establishes a national level accreditation that standardizes for the production, handling, and labeling of organically produced products. Included in these standards are lists of substances approved for and prohibited from use in organic production (see Appendix 1). For a more complete review of the NOP standards (Rule) the reader is urged to visit the NOP web site at http://www.ams.usda.gov/nop/. The portion of the NOP regulation pertaining to livestock operations is listed in Appendix 2. Dairy herds that desire to produce certified organic milk are a key component of the United States organic food production and have a significant presence in the NOP rulings. These rulings are very
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

130

comprehensive but can be reduce by saying that synthetics such as antibiotics, hormones, and pesticides are prohibited unless listed as approved. Natural products are acceptable unless listed as prohibited. Additionally, all feed must be purchased or produced as certified organic. Certified organic feed has to be grown on land free from application of manufactured fertilizer, pesticides, herbicides, and other chemicals for the prior 3 years. This certified organic feed must be fed to the herd and no prohibited substances may be used in cows producing certified organic milk. In addition to feed certification, product usage requirements are listed in the Rule pertaining to management systems and living conditions for the herd. This section in the Rule (205.239) is vague and has received a great deal of attention in organic dairy production as it pertains to pasture systems. Currently, the standards state that a producer of an organic livestock operation must establish and maintain livestock living conditions which accommodate the health and natural behavior of animals, including access to outdoors, shelter, shade, fresh air, and pasture (National Organic Program Standards). Controversy has surrounded this ruling especially as it pertains to pasture. On August 17th, 2005 the NOSB offered guidance for interpretation of the USDAs NOP pasture portion of The Organic System Plan (see Appendix 3). The NOSBs new recommendation states that ruminant livestock should graze pasture during the months of the year when pasture can provide edible forage. The Organic System Plan should have the goal of providing a significant portion of the total feed requirements as grazed feed but not less than 30% dry matter intake on an average daily basis during the growing season but not less than 120 days per year. Third party certification inspectors make thorough inspections of facilities, cows, and records at least annually to determine whether they conform to the NOP standards. Those producers who comply with the regulations are certified by a certifying agency and are allowed to use the USDA organic logo, make product statements, and sell their product as certified organic. Record keeping, accountability, and integrity are an integral part of managing a certified organic dairy within NOP guidelines. Additionally, key components of an organic herds management include a biosecurity plan, vaccination protocols, pasture plan, sanitation and hygiene routines, reproduction program, treatment and culling strategy, and well-balanced nutrition. It is important to understand that as stated in 205.238 (c)(7) of the NOP it is prohibited to withhold medical treatment from a sick animal in an effort to preserve its organic status. All appropriate medications must be used to restore an animal to health when methods acceptable to organic production fail. Livestock treated with a prohibited substance must be clearly identified and shall not be sold, labeled, or represented as organically produced. Feeding the organic dairy cow and nutritional principals associated with her wellness and productivity are no different than conventionally fed herds with one major exception. All feed ingredients must be certified organic or included in the NOP approved list of additives (see Appendix 1). The animal nutritional requirements, ration formulation, and least cost optimization are no different than those of conventional diets. Often the only challenge is procurement of organic certified grains and forages at an affordable cost. Conclusion The economics of organic dairy production are a fine balancing act between consumer demands creating a premium farm gate milk price, the higher prices for organic feeds, a lower feed efficiency (milk:feed ratio), and the added cost of production associated with the transition years. After a producer examines all the factors impacting profitability he or she will be in the position to choose the correct management system for their operation. If the decision is to become certified organic, then management of an organic dairy operation can be simply stated as; knowing the rules in the guidelines of the USDAs National Organic Program, staying within the boundaries of those rules even if you dont agree with them

131

or they dont make sense to you, and finally adhering to all best known management practices of dairy production management that modern science and technology offers. References

Martin, A. 2005. Organic Milk Debate. Chicago Tribune, January 10, Page 1. National Organic Program Standards, 2005. United States Department of Agriculture, Agricultural Marketing Service, 205.239 pp381-382.

Appendix 1 205.603 Synthetic substances allowed for use in organic livestock production.
In accordance with restrictions specified in this section the following synthetic substances may be used in organic livestock production: (a) As disinfectants, sanitizer, and medical treatments as applicable. (1) Alcohols. (i) Ethanol-disinfectant and sanitizer only, prohibited as a feed additive. (ii) Isopropanol-disinfectant only. (2) Aspirin-approved for health care use to reduce inflammation. (3) Biologics-Vaccines. (4) Chlorhexidine - Allowed for surgical procedures conducted by a veterinarian. Allowed for use as a teat dip when alternative germicidal agents and/or physical barriers have lost their effectiveness. (5) Chlorine materials - disinfecting and sanitizing facilities and equipment. Residual chlorine levels in the water shall not exceed the maximum residual disinfectant limit under the Safe Drinking Water Act. (i) Calcium hypochlorite. (ii) Chlorine dioxide. (iii) Sodium hypochlorite. (6) Electrolytes-without antibiotics. (7) Glucose. (8) Glycerin - Allowed as a livestock teat dip, must be produced through the hydrolysis of fats or oils. (9) Hydrogen peroxide. (10) Iodine. (11) Magnesium sulfate. (12) Oxytocin - use in post parturition therapeutic applications. (13) Paraciticides. Ivermectin - prohibited in slaughter stock, allowed in emergency treatment for dairy and breeder stock when organic system plan-approved preventive management does not prevent infestation. Milk or milk products from a treated animal cannot be labeled as provided for in subpart D of this part for 90 days following treatment. In breeder stock, treatment cannot occur during the last third of gestation if the progeny will be sold as organic and must not be used during the lactation period for breeding stock. (14) Phosphoric acid - allowed as an equipment cleaner, Provided, that, no direct contact with organically managed livestock or land occurs.

132

(b) As topical treatment, external parasiticide or local anesthetic as applicable. (1) Copper sulfate. (2) Iodine. (3) Lidocaine - as a local anesthetic. Use requires a withdrawal period of 90 days after administering to livestock intended for slaughter and 7 days after administering to dairy animals. (4) Lime, hydrated - as an external pest control, not permitted to cauterize physical alterations or deodorize animal wastes. (5) Mineral oil - for topical use and as a lubricant. (6) Procaine - as a local anesthetic, use requires a withdrawal period of 90 days after administering to livestock intended for slaughter and 7 days after administering to dairy animals. (c) As feed supplements - Milk replacers without antibiotics, as emergency use only, no nonmilk products or products from BST treated animals. (d) As feed additives. (1) DL - Methionine, DL-Methionine - hydroxy analog, and DL-Methionine - hydroxy analog calcium - for use only in organic poultry production until October 21, 2005. (2) Trace minerals, used for enrichment or fortification when FDA approved. (3) Vitamins, used for enrichment or fortification when FDA approved. (e) As synthetic inert ingredients as classified by the Environmental Protection Agency (EPA), for use with nonsynthetic substances or a synthetic substances listed in this section and used as an active pesticide ingredient in accordance with any limitations on the use of such substances. (1) EPA List 4 - Inerts of Minimal Concern. (f)-(z) [Reserved] [65 FR 80657, Dec. 21, 2000, as amended at 68 FR 61992, Oct. 31, 2003] 205.604 Nonsynthetic substances prohibited for use in organic livestock production. The following nonsynthetic substances may not be used in organic livestock production: (a) Strychnine. (b)-(z) [Reserved] 205.605 Nonagricultural (nonorganic) substances allowed as ingredients in or on processed products labeled as "organic" or "made with organic (specified ingredients or food groups(s))." The following nonagricultural substances may be used as ingredients in or on processed products labeled as "organic" or "made with organic (specified ingredients or food group(s))" only in accordance with any restrictions specified in this section. (a) Nonsynthetics allowed: Acids (Alginic; Citric - produced by microbial fermentation of carbohydrate substances; and Lactic). Agar-agar. Animal enzymes (Rennet - animals derived; Catalase bovine liver; Animal lipase; Pancreatin; Pepsin; and Trypsin). Bentonite. Calcium carbonate. Calcium chloride. Calcium sulfate - mined. Carageenan.

133

Colors, nonsynthetic sources only. Dairy cultures. Diatomaceous earth - food filtering aid only. Enzymes--must be derived from edible, nontoxic plants, nonpathogenic fungi, or nonpathogenic bacteria. Flavors, nonsynthetic sources only and must not be produced using synthetic solvents and carrier systems or any artificial preservative. Glucono delta-lactone production by the oxidation of D-glucose with bromine water is prohibited. Kaolin. Magnesium sulfate, nonsynthetic sources only. Nitrogen - oil-free grades. Oxygen--oil-free grades. Perlite--for use only as a filter aid in food processing. Potassium chloride. Potassium iodide. Sodium bicarbonate. Sodium carbonate. Tartaric acid. Waxes - nonsynthetic (Carnauba wax; and Wood resin). Yeast - nonsynthetic, growth on petrochemical substrate and sulfite waste liquor is prohibited (Autolysate; Bakers; Brewers; Nutritional; and Smoked - nonsynthetic smoke flavoring process must be documented). (b) Synthetics allowed: Alginates. Ammonium bicarbonate - for use only as a leavening agent. Ammonium carbonate - for use only as a leavening agent. Ascorbic acid. Calcium citrate. Calcium hydroxide. Calcium phosphates (monobasic, dibasic, and tribasic). Carbon dioxide. Cellulose - for use in regenerative casings, as an anti-caking agent (non-chlorine ableached) and filtering aid. Chlorine materials - disinfecting and sanitizing food contact surfaces, Except, that residual chlorine levels in the water shall not exceed the maximum residual disinfectant limit under the Safe Drinking Water Act (Calcium hypochlorite; Chlorine dioxide; and Sodium hypochlorite). Ethylene - allowed for postharvest ripening of tropical fruit and degreening of citrus. Ferrous sulfate - for iron enrichment or fortification of foods when required by regulation or recommended (independent organization). Glycerides (mono and di) - for use only in drum drying of food. Glycerin - produced by hydrolysis of fats and oils. Hydrogen peroxide. Lecithin - bleached. Magnesium carbonate - for use only in agricultural products labeled "made with organic (specified ingredients or food group(s))," prohibited in agricultural products labeled "organic." Magnesium chloride - derived from sea water. Magnesium stearate - for use only in agricultural products labeled "made with organic (specified ingredients or food group(s))," prohibited in agricultural products labeled "organic." Nutrient vitamins and minerals, in accordance with 21 CFR 104.20, Nutritional Quality

134

Guidelines For Foods. Ozone. Pectin (low-methoxy). Phosphoric acid - cleaning of food-contact surfaces and equipment only. Potassium acid tartrate. Potassium tartrate made from tartaric acid. Potassium carbonate. Potassium citrate. Potassium hydroxide - prohibited for use in lye peeling of fruits and vegetables except when used for peeling peaches during the Individually Quick Frozen (IQF) production process. Potassium iodide - for use only in agricultural products labeled "made with organic (specified ingredients or food group(s))," prohibited in agricultural products labeled "organic." Potassium phosphate - for use only in agricultural products labeled "made with organic (specified ingredients or food group(s))," prohibited in agricultural products labeled "organic." Silicon dioxide. Sodium citrate. Sodium hydroxide - prohibited for use in lye peeling of fruits and vegetables. Sodium phosphates - for use only in dairy foods. Sulfur dioxide - for use only in wine labeled "made with organic grapes," Provided, that, total sulfite concentration does not exceed 100 ppm. Tartaric acid. Tocopherols - derived from vegetable oil when rosemary extracts are not a suitable alternative. Xanthan gum. (c)-(z) [Reserved] [65 FR 80657, Dec. 21, 2000, as amended at 68 FR 61993, Oct. 31, 2003, and 68 FR 62217, Nov 3, 2003]

205.606 Nonorganically produced agricultural products allowed as ingredients in or on processed products labeled as "organic" or "made with organic (specified ingredients or food group(s))." The following nonorganically produced agricultural products may be used as ingredients in or on processed products labeled as "organic" or "made with organic (specified ingredients or food group(s))" only in accordance with any restrictions specified in this section. Any nonorganically produced agricultural product may be used in accordance with the restrictions specified in this section and when the product is not commercially available in organic form. (a) Cornstarch (native). (b) Gums - water extracted only (arabic, guar, locust bean, carob bean). (c) Kelp - for use only as a thickener and dietary supplement. (d) Lecithin - unbleached. (e) Pectin (high-methoxy).

Appendix 2 205.236 Origin of livestock.


(a) Livestock products that are to be sold, labeled, or represented as organic must be from livestock under continuous organic management from the last third of gestation or hatching: Except, That,

135

(1) Poultry. Poultry or edible poultry products must be from poultry that has been under continuous organic management beginning no later than the second day of life; (2) Dairy animals. Milk or milk products must be from animals that have been under continuous organic management beginning no later than 1 year prior to the production of the milk or milk products that are to be sold, labeled, or represented as organic, Except, That, when an entire, distinct herd is converted to organic production, the producer may: (i) For the first 9 months of the year, provide a minimum of 80-percent feed that is either organic or raised from land included in the organic system plan and managed in compliance with organic crop requirements; and (ii) Provide feed in compliance with 205.237 for the final 3 months. (iii) Once an entire, distinct herd has been converted to organic production, all dairy animals shall be under organic management from the last third of gestation. (3) Breeder stock. Livestock used as breeder stock may be brought from a nonorganic operation onto an organic operation at any time: Provided, That, if such livestock are gestating and the offspring are to be raised as organic livestock, the breeder stock must be brought onto the facility no later than the last third of gestation. (b) The following are prohibited: (1) Livestock or edible livestock products that are removed from an organic operation and subsequently managed on a nonorganic operation may be not sold, labeled, or represented as organically produced. (2) Breeder or dairy stock that has not been under continuous organic management since the last third of gestation may not be sold, labeled, or represented as organic slaughter stock. (c) The producer of an organic livestock operation must maintain records sufficient to preserve the identity of all organically managed animals and edible and nonedible animal products produced on the operation.

205.237 Livestock feed.


(a) The producer of an organic livestock operation must provide livestock with a total feed ration composed of agricultural products, including pasture and forage, that are organically produced and, if applicable, organically handled: Except, That, nonsynthetic substances and synthetic substances allowed under 205.603 may be used as feed additives and supplements. (b) The producer of an organic operation must not: (1) Use animal drugs, including hormones, to promote growth; (2) Provide feed supplements or additives in amounts above those needed for adequate nutrition and health maintenance for the species at its specific stage of life;

136

(3) Feed plastic pellets for roughage; (4) Feed formulas containing urea or manure; (5) Feed mammalian or poultry slaughter by-products to mammals or poultry; or (6) Use feed, feed additives, and feed supplements in violation of the Federal Food, Drug, and Cosmetic Act.

205.238 Livestock health care practice standard.


(a) The producer must establish and maintain preventive livestock health care practices, including: (1) Selection of species and types of livestock with regard to suitability for site-specific conditions and resistance to prevalent diseases and parasites; (2) Provision of a feed ration sufficient to meet nutritional requirements, including vitamins, minerals, protein and/or amino acids, fatty acids, energy sources, and fiber (ruminants); (3) Establishment of appropriate housing, pasture conditions, and sanitation practices to minimize the occurrence and spread of diseases and parasites; (4) Provision of conditions, which allow for exercise, freedom of movement, and reduction of stress appropriate to the species; (5) Performance of physical alterations as needed to promote the animal's welfare and in a manner that minimizes pain and stress; and (6) Administration of vaccines and other veterinary biologics. (b) When preventive practices and veterinary biologics are inadequate to prevent sickness, a producer may administer synthetic medications: Provided, that, such medications are allowed under 205.603. Parasiticides allowed under 205.603 may be used on (1) Breeder stock, when used prior to the last third of gestation but not during lactation for progeny that are to be sold, labeled, or represented as organically produced; and (2) Dairy stock, when used a minimum of 90 days prior to the production of milk or milk products that are to be sold, labeled, or represented as organic. (c) The producer of an organic livestock operation must not: (1) Sell, label, or represent as organic any animal or edible product derived from any animal treated with antibiotics, any substance that contains a synthetic substance not allowed under 205.603, or any substance that contains a nonsynthetic substance prohibited in 205.604. (2) Administer any animal drug, other than vaccinations, in the absence of illness; (3) Administer hormones for growth promotion;

137

(4) Administer synthetic parasiticides on a routine basis; (5) Administer synthetic parasiticides to slaughter stock; (6) Administer animal drugs in violation of the Federal Food, Drug, and Cosmetic Act; or (7) Withhold medical treatment from a sick animal in an effort to preserve its organic status. All appropriate medications must be used to restore an animal to health when methods acceptable to organic production fail. Livestock treated with a prohibited substance must be clearly identified and shall not be sold, labeled, or represented as organically produced. 205.239 Livestock living conditions. (a) The producer of an organic livestock operation must establish and maintain livestock living conditions, which accommodate the health and natural behavior of animals, including: (1) Access to the outdoors, shade, shelter, exercise areas, fresh air, and direct sunlight suitable to the species, its stage of production, the climate, and the environment; (2) Access to pasture for ruminants; (3) Appropriate clean, dry bedding. If the bedding is typically consumed by the animal species, it must comply with the feed requirements of 205.237; (4) Shelter designed to allow for: (i) Natural maintenance, comfort behaviors, and opportunity to exercise; (ii) Temperature level, ventilation, and air circulation suitable to the species; and (iii) Reduction of potential for livestock injury; (b) The producer of an organic livestock operation may provide temporary confinement for an animal because of: (1) Inclement weather; (2) The animal's stage of production; (3) Conditions under which the health, safety, or well being of the animal could be jeopardized; or (4) Risk to soil or water quality. (c) The producer of an organic livestock operation must manage manure in a manner that does not contribute to contamination of crops, soil, or water by plant nutrients, heavy metals, or pathogenic organisms and optimizes recycling of nutrients

138

Appendix 3
NOSB Livestock Committee Recommendation for Guidance Pasture Requirements for the National Organic Program March 2, 2005 Introduction The USDA National Organic Program (NOP) has requested NOSB provide guidance concerning the pasture requirements of the National Organic Program that the NOP can review and distribute to accredited certifying agents and post on the NOP website. The NOSB is seeking comments on organic system plan requirements; temporary confinement; and what constitutes appropriate pasture conditions. In particular, the NOSB seeks input on specific dry matter intake from pasture language; reference to regional NRCS prescribed grazing standards; and whether or not any of the text below should be recommended to the NOP for rule change. Guidance for interpretation of 205.239(a)(2) A. Organic System Plan Ruminant livestock shall graze pasture during the months of the year when pasture can provide edible forage. The Organic System Plan shall have the goal of providing grazed feed greater than 30% dry matter intake on a daily basis during the growing season but not less than 120 days. The Organic System Plan shall include a timeline showing how the producer will satisfy the goal to maximize the pasture component of total feed used in the farm system. For livestock operations with ruminant animals, the operations Organic System Plan shall describe: 1) The amount of pasture provided per animal; 2) The average amount of time that animals are grazed on a daily basis; 3) The portion of the total feed requirement that will be provided from pasture; 4) Circumstances under which animals will be temporarily confined; 5) The records that are maintained to demonstrate compliance with pasture requirements. B. Temporary Confinement Temporary confinement means the period of time when ruminant livestock are denied pasture. The length of temporary confinement will vary according to the conditions on which it is based (such as the duration of inclement weather) and instances of temporary confinement shall be the minimum time necessary. In no case shall temporary confinement be allowed as a continuous production system. All instances of temporary confinement shall be documented in the Organic System Plan and in records maintained by the operation. Temporary confinement is allowed only in the following situations: 1) During periods of inclement weather such as severe weather occurring over a period of a few days during the grazing season; 2) Conditions under which the health, safety, or well being of an individual animal could be jeopardized, including to restore the health of an individual animal or to prevent the spread of disease from an infected animal to other animals; 3) To protect soil or water quality C) Appropriate Pasture Conditions Appropriate pasture conditions shall be determined in accordance with the regional Natural Resources Conservation Service Conservation Practice Standards for Prescribed Grazing (Code 528) for the number of animals in the Organic Systems Plan.

139

NATIONAL RESOURCE AND CONSERVATION SERVICE DAIRY NUTRITION NUTRIENT MANAGEMENT RESEARCH PROJECTS IN THE MID-ATLANTIC REGION
Charlie Stallings, Rick Kohn, and Virginia Ishler Virginia Tech, Blacksburg, VA University of Maryland, College Park, MD Penn State, University Park, PA Phone: 540-231-3066 Fax: 540-231-5014 Email: cstallin@vt.edu Summary Funding agencies for animal nutrition conservation programs in the Mid-Atlantic region include Natural Resources Conservation Service (NRCS) and the National Fish and Wildlife Federation (NFWF). The NRCS funding is through Conservation Innovation Grants associated with the effort to reduce pollution of the Chesapeake Bay. The major emphasis is to use feed management to reduce overfeeding of protein and phosphorus on dairy farms. This can be done through improved ration formulation and feed delivery. The University of Maryland has taken the lead on using milk urea nitrogen to monitor overfeeding of protein. The Virginia emphasis is on phosphorus with an incentive program. Also in Virginia, ten herds are being precision fed using software to monitor ration delivery, consistency, and net flow of nutrients through the farm. The Pennsylvania study is looking at both nitrogen and phosphorus and use of precision feeding with feed analysis and adjustment of rations. All three states have projects three years in duration after which we will have much more information about how these feeding and management practices will be adopted. Current Programs in Feed Management from Maryland Conservation Innovation Grant NRCS funded Project Personnel: Rick Kohn, Telmo Oleas, University of Maryland Project Title: A program to improve dairy herd nutrition using milk urea nitrogen Project Duration: July 1, 2005 to June 30, 2008 The objectives are: 1) to institutionalize the routine measurement of milk urea nitrogen (MUN) on bulk-tank milk samples from three of the major milk cooperatives in the region, 2) to educate dairy farmers, educators (e.g. agricultural extension agents), technical assistance personnel (e.g. NRCS, soil conservation districts (SCD), private crop consultants) and representatives of allied industries (e.g. feed companies) about the use and interpretation of MUN results, 3) to identify dairy farms that have problems with herd nutrition, and provide them with needed assistance, 4) to demonstrate effectiveness of an incentive program designed to encourage nutritional consultants and dairy farmers to reduce nitrogen lost to the environment by decreasing nitrogen feeding, and 5) to integrate herd nutrition into comprehensive nutrient management plans. Dairy herd nutrition has a substantial impact on nutrient flows to water resources. Measurement of milk urea nitrogen (MUN) is an effective way to fine tune dairy herd diets and identify potential nutritional problems in herds. This project will provide an incentive for laboratories analyzing bulk-tank milk for farmer cooperatives to also analyze for milk urea nitrogen. Assistance will be given to the three largest cooperatives in Maryland to upgrade equipment and to measure MUN on a routine basis. Accuracy of sample analyses will be monitored by randomly testing milk from farms and comparing
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

140

results with those reported. Over 5000 farmers served by the cooperatives will receive information on interpreting MUN analyses and farms that have high MUN will be offered direct technical assistance. A one-time incentive program will be tested to encourage farmers to try reducing nitrogen in feed. Among 600 farmers that participate in the incentive program, those who are able to keep MUN below 11 mg/dl for 3 consecutive monthly averages will be awarded $150, and those who keep MUN monthly average below 12 mg/dl will be awarded $100. These levels are indicators of low nitrogen excretion and proper levels of protein in diets. After the program, cooperatives will be positioned to analyze MUN routinely on farms, and farmers and nutritional consultants will understand how to interpret the results. It is anticipated that farmers should be willing to pay the nominal cost for the analysis once the project is complete. At this time, all laboratories servicing cooperatives that handle milk in Maryland and Virginia have come on board and have upgraded their equipment. Currently we are working with laboratories to standardize procedures and quality control measures. Farmers have been informed about MUN results and interpretation. National Fish and Wildlife Federation (NFWF) Project Project personnel: Rick Kohn (UMD), Karin French (UMD), Virginia Ishler (Penn State), and Erica Cowan (Penn State) Project Title: Enhancing Nutrient Efficiencies on Dairy Farms in the Monocacy Watershed, MD and PA Objectives: Initiate precision dairy feeding on approximately 30 farms in the Monocacy watershed; Distribute educational materials, both print and audiovisual, on precision dairy feeding to over 300 dairy farmers, animal nutritionists, and feed industry representatives; In the short term (2-years), implementation of precision dairy feeding on 30 targeted farms will reduce the amount of nitrogen in manure by approximately 231,775 lb, based on a 21% reduction estimate; In 4 to 6 years, half of the watersheds dairy farmers will be using a precise feed management strategy.

The precision feeding program is part of a larger nutrient management program. A partnership of agricultural experts, including SCD, state agencies, university researchers, extension staff, and non-profit organizations propose to address nutrient pollution from dairy operations in the Monocacy River watershed in Maryland and Pennsylvania. This project will demonstrate that the comprehensive use of key management strategies could reduce nutrient losses to the environment by as much as 30-40%. Innovative, precise feed management strategies will be implemented on at least 20 farms within the watershed, reducing nitrogen and phosphorus overfeeding of dairy animals. Novel, farmer-friendly delivery systems will greatly increase the adoption of: (1) manure optimization on and between farms in the watershed to achieve a nutrient balance, and (2) enhanced small grain management and early planting of cover crops to trap residual nitrogen at the end of the growing season. Using results from a survey distributed to each farmer, analysis of feed, and milk analysis for MUN, we will identify dairy producers who are most likely to benefit by inclusion in the program. Once those producers have been identified, we will work directly with their nutritionist to improve feed management including: ration formulation, feed ingredient analysis (including dry matter and nutrient content), ration mixing, load cell accuracy, particle size analysis, evaluation of water quality, and evaluation of feed bunk space. The 300 dairy farmers in the Monocacy watershed have been contacted and all have been offered a free ration analysis by wet chemistry and evaluation of their diet formulation and feed delivery system. Virginia Ishler and Erica Cowan from Penn State have 8 out of the 11 viable farms in Pennsylvania located in the Monocacy watershed already committed to the project. They have done the preliminary

141

sampling for those herds and held their first educational meetings with the producers and their nutritionists. EQIP Cost Share Pilot Dairy farms identified in the above NFWF program will be referred to NRCS for potential for cost share payments to improve feed management. Payments will include up to $750 for nutritional management consultant, and $15 per cow (to a total of $4500/owner) for improved feed management to within 110% of NRC recommendations. Results will be verified with a combination of routine feed analysis and milk analysis for MUN. Applications are currently being accepted by NRCS for farmers in the Monocacy watershed only, as a pilot program for 2007. An expanded program is possible. Nutritionist Training and Certification Program Participants in the above EQIP and NFWF programs must agree to employ nutritionists who are certified to work with NRCS and University of Maryland on these programs. The certification training ensures the nutritionists understand the feed management evaluation process for NFWF program as the nutritionists will implement the improvements to feed management based on University recommendations. Participants will also be educated on feeding and reporting requirements for EQIP. Current programs in feed management from Virginia Conservation Innovation Grant NRCS and Virginia Department of Conservation and Recreation funded Project Title: Virginia Precision Phosphorus Feeding Incentive Program Project personnel: Charlie Stallings, Katharine Knowlton, Bob James, and Mark Hanigan, Virginia Tech Project Duration: Fall 2005 through 2009 Research reports with large numbers of high producing cows and across multiple lactations have revealed the desirability of feeding phosphorus at lower levels than has been reported from a survey of Virginia dairy herds. The 2001 National Research Council (NRC) supports a lower supplementation rate. The Department of Dairy Science at Virginia Tech is conducting a voluntary incentive program. All Grade A permitted herds received an introductory survey in October 2005 asking about the interest in participating. Selected herd owners that indicated interest in the incentive program were contacted and signed up in four groups staggered so that each group started at a different time of the year in 2006. At the end of 2006 there were 183 Virginia dairy herds participating in the project. Below are details about the project. Reasons a dairy might participate: 1. Prepare farm for pending rule changes. Starting January 1, 2007 phosphorus based plans are needed in some operations where soil phosphorus and erodability is an issue (P Index). 2. Even if Number 1 does not apply, this project will document the dairy industries willingness to reduce nutrient output and benchmark the level of supplementation. 3. Free feed testing for phosphorus and major nutrients. 4. Free ration consultation is available if requested. 5. If goals for phosphorus intake are reached a payment can be made. Expectations of the dairy farm: 1. The herd owner/manager will meet with project personnel at the beginning of the project and discuss expectations. Information on feeds fed and amounts, milk production, number of cows, fat test, and breed will be collected.

142

2. Total mixed rations and other feeds, up to three feeds per farm per sampling period, will be sampled according to established protocol and sent by the owner/manager every other month to Cumberland Valley Analytical Services. Postage will be paid by the project. Any change in feeds fed or amounts as well as milk production and cow numbers should be reported at this time. 3. Once a year, project personnel will be allowed to make a check visit and obtain test samples and herd information for verification. 4. Payments can be received at the end of the first and second year but feed sample collection and analysis will continue for a third year to document long term impact. Reporting Procedures and calculation of payments: 1. All data collected will be confidential and individual results will only be available to the farmer and project personnel. Summaries may be made with some data but only with averages from five or more farms. 2. Cumberland Valley Analytical Services will report results of the lab analysis, calculated intake of phosphorus, required phosphorus amount, and intake as a percent of required. 3. At the end of the first and second year a standard deviation from the six yearly measurements will be calculated and any unusual sample (one that is above or below the average the standard deviation) will be discarded. 4. If the check test does not match previous results, project personnel will collect another check sample. If further tests reveal differences it will be at the discretion of project personnel to determine payment eligibility. 5. There will be two-tier compensation with herds feeding phosphorus at less than 105% of 2001 NRC requirements receiving $12 per cow per year and herds feeding between 105 and 115% of requirements receiving $6 per cow per year. In some situations a farm might qualify for one year but not the other. Herds above 115% excess will receive no direct payment but will receive the free feed testing and ration consultation. 6. Maximum payment per farm will be $4,800 for one year or $9,600 for the two-year period. Other information related to the project: This project is an educational effort to foster awareness of an issue important in Virginia and other states. It is for herds being overfed phosphorus as well as those that have already modified their feeding practices. Educational materials have been prepared periodically dealing with related issues and assistance to the herd nutritionist is provided upon request. There may be situations where it is not cost effective to formulate diets to meet the required ranges for payment. It is up to the individual farm and their consultants to determine economic feasibility. Factors such as geographical location and complexity of the feeding program may be considerations for enrollment. For further information contact the Virginia Tech Dairy Science Extension web site address: http://www.vtdairy.dasc.vt.edu/. Ten herds were selected to be part of an intensive part of this project where feed management software was installed on each farm for inventory control and feeder performance evaluation. These herds will have nutrient flow calculations made for their farm to determine net import or export of nitrogen and phosphorus. Feed sampling on all feeds is done monthly. First year results will be available in 2007. Current programs in Feed Management from Pennsylvania Conservation Innovation Grant NRCS funded Title: Precision Dairy Feeding to Reduce Nutrient Pollution in Pennsylvanias Waters and the Chesapeake Bay Project personnel: Kelly ONeill and Matthew Ehrhart, Chesapeake Bay Foundation, Dr. Robert Munson and Dr. James Ferguson, University of Pennsylvania, Barry Frantz and Gary Smith, USDA Natural Resources Conservation Service, and Virginia Ishler and Tim Beck, Pennsylvania State University

143

Project duration: July 1, 2005 through June 30, 2008 Almost 4,000 miles of Pennsylvania streams are impaired by agricultural non-point source pollution, primarily due to excessive loadings of sediment and nutrients. Based on livestock numbers and densities, the primary source of those nutrients is livestock manure. Utilizing the best available data and phosphorus-based nutrient management plans; the NRCS has calculated excess manure for each county in the U.S. The Lower Susquehanna basin alone generates 286,196 tons of excess manure. The Chesapeake 2000 Agreement established commitments to be met by 2010 in order to defer the construction of a Bay watershed-wide Total Maximum Daily Load (TMDL). Implementing precision dairy feeding practices broadly is one of the keys to meeting the objectives of local water quality based TMDLs, as well as Chesapeake Bay water quality commitments. The scientific body of evidence accumulated by the efforts to restore the Chesapeake Bay has clearly established the need to resolve this overload of manure nutrients. Pennsylvanias agricultural nonpoint source driven TMDLs recognize the impact of excess manure on the quality of rivers and streams. Solutions for dealing with excess dairy manure are difficult due to the liquid nature of the waste and the diffuse nature of its production. A significant percentage of the dairy manure is termed unrecoverable because it is deposited directly on pastureland by unconfined livestock. One of the most effective ways to reduce nitrogen pollution from dairy manure is to manipulate dietary formulations to meet herd nutritional requirements with less dietary nitrogen. Objectives: Pennsylvanias Tributary Strategy indicates that precision dairy feeding will need to be implemented in 75% of the states dairy cows by 2010 to meet the States commitment to the Chesapeake 2000 Agreement. This precision dairy feeding program will begin that transformation and set the stage for the broader industry adoption of this critical practice. Dairy farmers traditionally have incorporated excess nitrogen and phosphorus in the rations to ensure health and maximize production. This made far more sense under historic conditions, when livestocks contribution to impairment of surface water was less known and the dairy industry was not facing pressure to find alternatives for manure management. Moreover, scientific research is now available to more accurately assess nutrient needs of livestock, as well as the nutrient content of the various feeds available. This project will provide the necessary resources to assist dairy producers in examining forage quality, ration development, and monitoring for nutrient deficiencies. Participating operations should see at least a 25% reduction in nitrogen and phosphorus in manure, with positive impacts on profitability, manure nutrient management burdens and water quality. Project Methods: Numerous hurdles have impeded the implementation of precision feeding to date. In part, this is due the diversity of the dairy industry because of the variation in size of operations and management styles of dairy producers. Secondly, the number of feed company representatives, private nutritionists and veterinarians, each with varying degrees of nutritional knowledge make implementation of uniform industry-wide changes difficult. The goal of nutritionists and feed consultants has always been to maximize milk production, with little or no concern for the impact of excess nutrients. Most dairy producers do not have a background in nutrition and rely on the advice of others, not knowing or being able to estimate the costs of excess feed nutrients. The information and decision making skills to evaluate precision feeding have not been in the hands of producers or many of those advising producers on nutrition issues. Other hurdles to the desired industry shift are labor and time. For all dairies, small dairies in particular, the process and practices necessary to implement precision feeding must be simplified and 144

integrated into the day-to-day farm operations in an efficient fashion. Dairy farms rely on forages that have wide variations in quality, digestibility and nutrient content that are the result of variation in soil type, weather at harvest, and many other circumstances beyond the producers control. The implementation process must address not only the changes that need to be made, but how those changes can be seamlessly integrated into farm operations. The expertise of the project partners, and others involved with the Pennsylvania Secretary of Agricultures Dairy Task Force, will be utilized to surmount the existing hurdles and demonstrate the value and efficiencies of precision dairy feeding. The Chesapeake Bay Foundation, in partnership with veterinarians from the University of Pennsylvanias New Bolton Center, Pennsylvania State University Cooperative Extension Service, the Natural Resources Conservation Service (NRCS), and the Pennsylvania Department of Agriculture is pursuing ambitious goals to bring about significant changes in the dairy industrys standard feeding practices and subsequently improve water quality. This projects three-year goals under this grant are as follows: Initiate precision dairy feeding on 60 farms that will receive cost-share assistance to cover the necessary laboratory analyses, technical assistance in interpreting data and adjusting rations and management. To date, 40 dairy farms are participating in the project. Develop and distribute educational materials on precision dairy feeding to over 3,000 Pennsylvania dairy producers. Educational materials have been developed and provided on a CD during the educational workshops. They are also available on the web at

http://www.das.psu.edu/dairynutrition/
Develop and conduct 12 workshops on precision dairy feeding to reach 300 veterinarians, animal nutritionists and feed industry representatives. To date, 8 workshops have been held and four more are scheduled for the winter of 2007. One hundred thirty-six consultants have participated in the educational program: Profitability Assessment Dairy Tool, with 114 going through the drill downs to precision feeding.

The project partners will implement a program to obtain the necessary farm information, conduct the relevant analyses, design the appropriate feed ration, and implement that ration on Pennsylvania dairy farms. Each participating farm will receive intensive technical assistance to ensure that herd health and production are maintained or improved. The program will also include a comprehensive education strategy to reach farmers, animal nutritionists, veterinarians, and livestock feed industry staff. The Chesapeake Bay Foundation will work with its partners to deliver the grant commitments, utilizing the unique expertise and relationships of each entity. The University of Pennsylvanias New Bolton Center has been contracted to provide the necessary on farm sampling and data acquisition, ration formulation, and technical expertise. Collected data will include: forage analysis, total mixed ration (TMR) analysis, fecal and urine analysis, and MUN analysis. Technicians from the New Bolton Center will collect samples and data from the selected farms. New Bolton staff will utilize this information to formulate the appropriate feed ration for the farms dairy herd. The feed rations will be designed using CPM (Cornell Penn Miner), a model that has been used throughout the world for many years to accurately formulate rations and balance amino acids for dairy cows throughout their lactation cycle. The project partners will work with the dairy operation to implement the strategies and procedures in their operation to ensure that the formulated ration is reaching the animals, including the use of MUN analysis. Followup data collection of manure and urine will document the reduction in nutrient output from the dairy herd. The Chesapeake Bay Foundation is working with the Pennsylvania State University Cooperative Extension Service to produce and distribute the educational products that are critical to broader implementation of this practice. Materials will include a brochure for dairy farms outlining the details and advantages of precision feeding and a web site to provide instruction for farmers and farm workers on how to implement the process with their dairy herd.

145

The ultimate goal of this program is to catalyze a major industry shift to precision feeding in order to reduce nitrogen in Pennsylvanias dairy manure by tens of millions of pounds. References Dou, Z., R.A. Kohn, J.D. Ferguson, R.C. Boston, and J.D. Newbold. 1996. Managing nitrogen on dairy farms: An integrated approach I. Model Description. J. Dairy Sci. 79:2071-2080. Knowlton, K.F., and J.H. Herbein. 2002. Phosphorus partitioning during early lactation in dairy cows fed diets varying in phosphorus content. J. Dairy Sci. 85:1227-1236. Sniffen, C.J., and W. Chalupa. 2003. The Lactating Dairy Cow and Its Amino Acid Needs. Feed Management, July/August Volume 54, No. 4. http://www.jefo.ca/pdf/Ruminants/dairy%20amino%20acid%20needs.pdf.

146

CHALLENGES IN DETERMINING ENERGY REQUIREMENTS OF HORSES


Laurie Lawrence Department of Animal and Food Sciences University of Kentucky 905 Garrigus Lexington KY 40546-0215 Phone: 859-257-7509 FAX: 859-257-2534 Email: llawrenc@uky,edu Summary Energy requirements for horses have been expressed as megacalories of digestible energy (DE) in the National Research Council (NRC) nutritional requirement publications since 1961. Net energy systems are currently used for beef and dairy cattle. A net energy system for horses has been studied in France. Net energy systems provide the most detailed explanation of energy catabolism but they are complex and require some information not readily available for horses. The DE system is less detailed but also less complicated and thus is easier to apply in many practical situations. The daily DE requirements of horses may be partitioned into two general categories: maintenance energy and productive energy. The latter category is the DE that is available for weight gain, milk production, fetal tissue accumulation or work. For most horses, more energy is expended on maintenance each day than on a productive purpose. Therefore, accurate estimation of the maintenance component of the daily energy requirement is very important. It is also very difficult. The domestic horse population is very diverse. The maintenance requirement of an individual horse may be affected by age, mature body size, temperament, body composition and housing conditions. The 6th edition of the Nutrient Requirements of Horses (NRC, 2007) has attempted to recognize this diversity by providing a range of maintenance levels. It also provides information on the factors that induce variability in energy requirements of all classes of horses. DE, ME, NE: Which Energy System? Optimal rations for horses match energy intake to energy need. Therefore, an effective energy system is one that can be applied to both determining the useful energy in feed and determining the energy requirement the animal. The Nutrient Requirements of Horses was initially published in 1961 as a revision of a previous report entitled Recommended Nutrient Allowances for Horses (NRC, 1949). In the original report, energy requirements were not specifically defined, but daily allowances for total digestible nutrients (TDN) were given. Since 1961, the energy requirements of horses (NRC, 1966; 1973; 1978; 1989) have been expressed in units of digestible energy (DE). The DE system is relatively simple to use because information on the DE content of many feeds is available. However, it is not a very precise system. In cattle, energy requirements are often expressed in units of net energy (NRC, 2000; 2001). In the net energy systems available for beef and dairy cattle, the net energy values of feeds are related to their biological use. Therefore, a single feed can have one net energy value if it is used primarily for maintenance and other values when it is used primarily for growth or lactation. A net energy system for horses has been described by researchers in France (Martin-Rosset et al., 1994; Martin-Rosset, 2000; Martin-Rosset and Ellis, 2005; Vermorel and Martin-Rosset, 2004. This system has been applied in a number of research studies, particularly in regard to maintenance horses (Martin-Rosset and Vermorel, 1991; Vermorel et al., 1991; Vermorel et al., 1997a, 1997b). An advantage of the French net energy system is that it allows for adjustments to energy requirements based on the efficiency with which various energy sources (for example, fatty acids versus glucose) are used. A disadvantage of the French net energy system is that it expresses energy requirements in horse feed units rather than calories or joules.

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

147

The French horse feed unit is based on barley, and all other feeds are evaluated in relation to the net energy in barley. In the 6th edition of the Nutrient Requirements of Horses (NRC, 2007), the DE system has been retained as the means of expressing energy requirements. The advantages of a net energy system are acknowledged in the document, and it should be noted that a net energy approach was used as a step in the process of estimating energy requirements for some classes of horses. However, because the net energy values of many feeds were lacking, the NE system was not used to define daily dietary energy intakes. Equus: Consider the Diversity In other livestock species, many selection criteria have been related to variables such as composition of gain, milk production or feed efficiency, thus reducing the variation among the population in regard to nutrient requirements. These characteristics have not been used for selection in horses. Instead, horses have been selected for characteristics related to their athletic function. Some breeds have been selected for physiological characteristics that facilitate speed over short distances, while others have been selected for endurance activities. Horses have also been selected for characteristics related to temperament and suitability for use in various activities that require a high level of trainability. Domestic equids range in body weight (BW) from less than 200 kg to more than 1000 kg. Horses can live for 20 years or more, making them long-lived compared to other livestock. Little is known about how these differences in physiological characteristics, temperament, BW or age affect nutrient requirements of horses. Thus, extrapolating energy requirements determined in a group of horses with one set of characteristics (age, BW, breed, gender) to other groups of horses is fraught with pitfalls. For example, most studies in North America on growing horses have used Thoroughbreds or Quarter Horses (Coleman et al., 1997; Cymbaluk et al., 1989; Cymbaluk and Christison, 1989; Cymbaluk, 1990; Lawrence et al., 1991; Ott and Asquith, 1986, 1995; Ott and Kivipelto, 1998, 1999; Ott et al., 1979, 2005; Pagan et al., 1996, 2005; Scott et al., 1989; Staniar et al., 2004). Only a few studies have used other popular light horse breeds such as Arabians, Morgans, Saddlebreds, Tennessee Walking Horses or Standardbreds (Bell et al., 2001; Lawrence et al., 2003; Reed and Dunn, 1977; Sandgren et al., 1993). Some of the studies have used ponies (Ousey et al., 1997) but studies on the growth requirements of modern draft breeds in the U.S. are essentially non-existent. Despite the absence of specific information for horses of all breeds and mature BW, The Nutrient Requirements of Horses (NRC, 2007) provides estimates of nutrient requirements for horses with mature BW between 200 and 900 kg. Because most of the recommendations in the NRC (2007) are drawn from data generated using breeds with mature BW of 500 to 600 kg, the estimates for much larger and much smaller horses may lack precision. Estimating Maintenance Horses that are not growing, lactating, gestating or working are often considered to be at maintenance. However, the maintenance requirement is a component of the daily energy requirement for all classes of horses. For example, the daily energy requirement for a growing horse is the sum of the energy required for tissue accretion (weight gain) and the energy required to maintain the existing tissues in the horses body. The maintenance component comprises more than 50% of the daily energy requirement for almost all horses. For many horses, maintenance accounts for almost all the daily requirement. Therefore, to accurately estimate daily energy requirements, one must be able to accurately estimate the maintenance component. This is not easy; the maintenance component can be influenced by many factors. Maintenance requirements are sometimes expressed on the basis of an adjusted body weight (metabolic body weight). A metabolic body weight is employed because maintenance requirements do not vary in direct proportion to body weight when animals of vastly different body weights are considered. Prior to 1989, the NRC (1978) used metabolic body weight (BW0.75) to calculate energy requirements for

148

horses. Using the NRC (1978) method the maintenance DE for a 200-kg pony was approximately 41 kcal/kg BW, compared to 27 kcal/kg BW for a 900-kg draft horse. In 1989, DE requirements for maintenance were calculated using actual body weight. However, the equations that were used still resulted in higher DE allowances per kg BW for small animals compared to large animals (37 kcal/kg BW for a pony; 27 kcal/kg BW for a draft horse). In the 6th edition of the publication (NRC, 2007), maintenance requirements are calculated on a direct BW basis, so that the requirements for horses of all mature BW are the same when expressed per kg BW. Other nutrient requirements are also expressed on the basis of actual BW, thus by expressing DE requirements on a BW basis, the nutrient:calorie ratios remain consistent across mature body sizes. It seems reasonable that a small animal with a higher surface area to mass ratio would have a relatively higher maintenance requirement than a large animal. However, ponies are often described as easy keepers in comparison to some larger breeds such as thoroughbreds. It seems contradictory for an easy keeping pony to have a higher maintenance requirement per kg BW than a hard keeping thoroughbred. Researchers have measured the maintenance requirements of horses using a number of techniques. A relatively simple and direct method is to weigh mature horses that are not working, lactating, or gestating and measure the amount of feed needed to maintain body weight. The DE in the feed can then be determined using a digestion trial and the DE for maintenance calculated. Alternatively, indirect calorimetry may be used to estimate resting heat production, and then daily dietary energy requirements may be calculated. The latter method requires information on the DE content of feeds, and the efficiency with which absorbed energy is converted to usable energy in the horse. This information may be obtained from metabolic balance studies. A review of studies on the maintenance requirements of horses is included in the NRC (2007). The estimated DE requirements for maintenance determined in indirect calorimetry/metabolic balance studies were generally lower than values determined in feeding studies (Anderson et al., 1983; Barth et al., 1977; Pagan et al., 1986; Potter et al., 1987; Stillions and Nelson, 1972; Vermorel et al., 1991; Vermorel et al., 1997a, 1997b; Wooden et al., 1970). Animals in balance or indirect calorimetry studies may be more confined than animals in feeding studies. In addition, they may have been maintained in more thermoneutral environments. Thus, maintenance values derived in balance or calorimetry studies probably underestimate the maintenance needs of some animals kept in practical environments. Alternatively, they may provide insight into the minimum amount of energy required by an animal with little activity or environmental stress. Sources of variation in an animals maintenance requirement can include age, body composition, the environment in which the animal lives and its diet. Horses in early adult-hood (approximately 4 years old) had higher maintenance requirements than middle-aged horses (approximately 11 years) (Vermorel and Martin-Rosset, 1991). The effect of body composition on maintenance energy requirements has not been measured in horses, but Birnie et al. (2000) reported that dairy cows with low condition scores had higher fasting heat production than fatter cows. Environments that stimulate thermoregulatory mechanisms can also affect the maintenance requirement. Fasting heat production is elevated when horses are exposed to temperatures below their lower critical temperature (McBride et al., 1985). Not all energy yielding feed components are metabolized with similar efficiency. Replacing carbohydrates with fat in the diet can reduce the total amount of DE needed by horses (Potter et al., 1989). Compared to the DE in fiber, the DE in fat and starch is metabolized more efficiently by horses; thus, the daily DE requirement would be expected to be higher on high fiber diets compared to low fiber diets. From the previous discussion it should be clear that there are many factors to consider when determining an individual horses maintenance requirement. Horsemen have recognized the variation in maintenance requirements with the terms easy keeper and hard keeper. Prior to 2007, the Nutrient Requirements of Horses did not specifically address this variation. The 6th edition of the publication presents three levels of maintenance for adult horses. The minimum maintenance value was obtained from a summary of studies that measured maintenance requirements in horses using indirect calorimetry and balance trials. Horses in the minimum maintenance category are likely to be horses in mid to late

149

maturity that are generally inactive. One could envision that these horses might fit the horsemens description of an easy keeper and tend towards higher condition scores. The average maintenance requirement was developed from previous recommendations as well as interpretation of feeding studies. For a 500 kg horse, the new (NRC, 2007) daily DE requirement is very similar to previous (NRC, 1978, 1989) recommendations. The average horse is likely to be in early to middle maturity and to have moderate voluntary activity. The third category is the elevated maintenance requirement. This category is designed for the hard keeper; including horses with high levels of voluntary activity. Horses in early maturity or horses that are easily stimulated by activity in their environments might be included in this category. The elevated maintenance value is used to calculate the maintenance component of the daily DE requirement of stallions and some lactating mares. As noted above, previous publications suggested that the maintenance requirement of a 200 kg mature pony on a weight basis (kcal/kg BW) was higher than the requirement for a 500 kg mature horse, which was higher than the requirement for a 900 kg draft horse. The new NRC (2007) does not make this assumption. As a result, the total daily DE intake for an average 200 kg mature pony (Mcal/d) listed in the new NRC (2007) is lower than the estimate given in the 1989 NRC. Further, the value for an average 900 kg mature draft horse is higher in the new NRC than in the previous edition. However, the new publication allows the user to determine whether the requirements of an active pony are better met by the elevated category or if the needs of a docile draft horse are better met by the minimum value. The new NRC (2007) indicates that these levels of maintenance should not be expected to fit every horse and that some individuals may have a maintenance requirement outside of the range defined by the minimum and elevated values. Gestation, Lactation, Growth and Exercise The energy requirement of a gestating mare is the sum of the energy needed to maintain the mares body and the energy needed for fetal, placental, uterine and mammary development. Energy deposition in the products of conception has not been carefully studied in horses. The only available data on the developing equine fetus are from studies on aborted foals (Giussani et al., 2005; Meyer and Ahlswede, 1978; Platt, 1978). Careful study of this area would require slaughter of pregnant mares at different stages of gestation. Based on the limited data in horses and some studies with other species (Bell et al., 1995; Ji et al., 2005; Reynolds et al., 1986) estimates of fetal, uterine and placental development were used to predict the energy needs of gestating mares above maintenance. These estimates were then added to the maintenance requirements to arrive at a daily nutrient intake. Most fetal development occurs in the last trimester of gestation (Meyer and Ahlswede, 1978; Platt, 1978). However, in other animals significant placental development and uterine hypertrophy precedes fetal development (Ji et al., 2005). Therefore, energy requirements may increase above maintenance before the 9th month of gestation. The new NRC (2007) includes energy increments above maintenance beginning in the 5th month of gestation. The energy requirement for lactating mares in the new NRC (2007) was derived with a method similar to the previous publication. Studies that measured milk composition and milk yield were summarized to produce estimates of daily milk energy secretion. The DE needed for milk production was then added to the maintenance requirement to obtain the total daily DE requirement. As noted previously, the elevated maintenance requirement was used for lactating mares. The elevated requirement was used because lactating mares have a higher feed intake (and thus higher maintenance heat production) and because maternal behavior may increase their daily voluntary activity level. The energy requirements for growing horses can be calculated by adding the energy needed for weight gain to the energy needed for maintenance (NRC, 1989). Carcass data has been used to obtain information on the composition of gain and the amount of energy retained by the growing food animals. Little is known about the composition of gain in growing horses, or whether rate of gain, breed type, or gender affect composition. Without this information the estimates of the amount of energy needed per unit of gain are not precise.

150

As with the other physiological classes, the requirement for an exercising horse is the sum of the energy needed for maintenance plus the energy needed for imposed exercise. Quantifying the amount of energy expended during exercise is difficult. Many factors affect energy use during exercise such as the speed and duration of exercise, the fitness of the animal, the amount of weight carried by the horse, the condition of the ground surface, etc. Heart rate has been proposed as a practical means of estimating exercise intensity and oxygen utilization which can then be used to estimate energy use (Coenen, 2005). A component of the daily energy requirement that is often overlooked is the effect of transport on horses. Many equine athletes will travel several hours to and from competition. At least one study suggested that the muscular effort associated with trailer transport is approximately the same as walking (Doherty et al., 1997). References Anderson, C.E., G.D. Potter, J.L. Krieder and C.C. Courtney. 1983. Digestible energy requirements of exercising horses. J. Anim. Sci. 56:91-95. Barth, K.M., J.W. Williams and D.G. Brown. 1977. Digestible energy requirements of working and nonworking ponies. J. Anim. Sci. 44:585-589. Bell, A.W., R. Sleptis and R.A. Ehrnhardt. 1995. Growth and accretion of energy and protein in the gravid uterus during late pregnancy in Holstein cows. J. Dairy Sci. 78:1954-1961 Bell, R.A., B.D. Nielsen, K.Waite, D. Rosenstein and M. Orth. 2001. Daily access to pasture turnout prevents loss of mineral in the third metacarpus of Arabian weanlings. J. Anim. Sci. 79:1142-1150. Birnie, J.W., R.E. Agnew and F. J. Gordon. 2000. The influence of body condition in the fasting energy metabolism of nonpregnant, nonlactating dairy cows. J. Dairy Sci. 83:1217-1223. Coenen, M. 2005. About the predictability of oxygen utilization and energy expenditure in the exercising horse. Page 123. In: Proc. Equine Science Soc. Symp, Tucson, AZ. Coleman, R.J., G.W. Mathison, L. Burwash and J.D. Milligan. 1997. The effect of protein supplementation of alfalfa diets on the growth of weanling horses. Pages 59-64. In: Proc. 15th Equine Nutr. Physiol. Symp., Ft. Worth TX. Cymbaluk, N.F., and G.I. Christison. 1989. Effects of diet and climate on growing horses. J. Anim. Sci. 67:48-59. Cymbaluk, N.F., G.I. Christison and D.H. Leach. 1989. Energy uptake and utilization by limit and ad libitum fed growing horses. J. Anim. Sci. 67:403-413. Cymbaluk, N.F. 1990. Cold housing effects on growth and nutrient demand of young horses. J. Anim. Sci. 68:3152-3162. Doherty, O., M. Booth, N. Waran, C. Salthouse and D. Cuddeford. 1997. Study of the heart rate and energy expenditure of ponies during transport. Vet. Rec. 141:589-592. Giussani, D.A., A.J. Forehead and A.L. Fowden. 2005. Development of cardiovascular function in the horse fetus. J. Physiol. 565:1019-1030. Ji, F., G. Wu, J.R. Blanton and S.W. Kim. 2005. Changes in weight and composition in various tissues of pregnant gilts and their nutritional implications. J. Anim. Sci. 83:366-375. Lawrence, L.A., H. R. Hearne, S.P. Davis, J.D. Pagan, A. Fitzgerald and E.A. Greene. 2003. Characteristics of growth in Morgan horses. Pages 317-322. In: Proc. 18th Equine Nutr. Physiol. Symp., East Lansing, MI. Lawrence, L., M. Murphy, K. Bump, D. Weston and J. Key. 1991. Growth responses in hand-reared and naturally reared Quarter Horse Foals. Equine Pract. 13:19. Martin-Rosset, W. 2000. Feeding standards for energy and protein for horses in France. Pages 31-94. In: Proc. 2000 Equine Nutrition Conference for Feed Manufacturers. Kentucky Equine Research, Versailles, KY. Martin-Rosset, W., and A.D. Ellis. 2005. Evaluation of energy and protein requirements and recommended allowances in growing horses. Pages 103-136. In: The Growing Horse: Nutrition and Prevention of Growth Disorders. EAAP Publication No. 114.

151

Martin-Rosset, W., and M. Vermorel. 1991. Maintenance energy requirement variations determined by indirect calorimetry and feeding trials in light horses. J. Equine Vet. Sci. 11:42-45. Martin-Rosset, W., and M. Vermorel. 2004. Evaluation and expression of energy allowances and energy value of feeds in the UFC system for the performance horse. Pages 29-60. In: Nutrition of the Performance Horse, EAAP Publication No. 111, Dijon, France. Martin-Rosset, W., M. Vermorel, M. Doreau, J.L. Tisserand and J. Andieu. 1994. The French horse feed evaluation systems and recommended allowances for energy and protein. Livestock Prod. Sci. 40:3756 McBride, G.E., R.J. Chistopherson and W. Sauer. 1985. Metabolic rate and plasma thyroid concentrations of mature horses in response to changes in ambient temperatures. Can. J. Anim. Sci. 65:375-382. Meyer, H. and L. Ahlswede. 1978. The intra-uterine growth and body composition of foals and the nutrient requirements of pregnant mares. Anim. Res. Devel. 8:86-111. NRC. (National Research Council). 1949. Recommended Nutrient Allowances for Horses. National Academy Press, Washington, D.C. NRC. (National Research Council). 1961. Nutrient Requirements of Horses. National Academy Press, Washington, D.C. NRC. (National Research Council). 1966. Nutrient Requirements of Horses. National Academy Press, Washington, D.C. NRC. (National Research Council). 1973. Nutrient Requirements of Horses. National Academy Press. Washington, D.C NRC (National Research Council). 1978. Nutrient Requirements of Horses. National Academy Press. Washington, D.C NRC (National Research Council). 1989. Nutrient Requirements of Horses. National Academy Press. Washington, D.C NRC (National Research Council). 2000. Nutrient Requirements of Beef Cattle. National Academy Press. Washington, D.C NRC (National Research Council). 2001. Nutrient Requirements of Dairy Cattle. National Academy Press. Washington, D.C NRC. (National Research Council). 2007. Nutrient Requirements of Horses. National Academy Press. Washington, D.C Ousey, J.C., S. Prandi, J. Zimmer, N. Holdstock and P.D. Rossdale. 1997. Effects of various feeding regimens on the energy balance of equine neonates. Amer. J Vet. Res. 58:1243-1251 Ott, E.A., and R.L. Asquith. 1986. Influence of level of feeding and nutrient content of the concentrate on the growth and development of yearling horses. J. Anim. Sci. 62:290-299. Ott, E.A, and R.L. Asquith. 1995. Trace mineral supplementation of yearling horses. J. Anim. Sci. 73:466. Ott, E.A., R.L. Asquith, J.P Feaster and F.G. Martin. 1979. Influence of protein level and quality on the growth and development of yearling foals. J. Anim. Sci. 49:620-628. Ott, E.A., M.P. Brown, G.D. Roberts and J.Kivipelto. 2005. Influence of starch intake on growth and skeletal development of weanling horses. J. Anim. Sci. 83:1033-1043. Ott, E.A., and J. Kivipelto. 1998. Influence of dietary fat and time of hay feeding on growth and development of yearling horses. J. Eq. Vet. Sci. 18:254-259. Ott, E., and J. Kivipelto. 1999. Influence of chromium tripicolinate on growth and glucose metabolism in yearling horses. J. Anim. Sci. 77:3022-3029. Pagan, J.D., and H.F. Hintz. 1986. Equine Energetics I. J. Anim. Sci. 63:816-822. Pagan, J., S. Jackson and S. Caddel. 1996. A summary of growth rates in thoroughbreds in Kentucky. Pferdeheilkunde 12(3):285 Pagan, J.D., A. Koch, S. Caddel and D. Nash. 2005. Size of thoroughbred yearlings presented for auction at Keeneland Sales affects selling price. Pages 224-225. In: Proc. 19th Equine Science Symp., Tuscon, AZ. Platt, H. 1978. Growth and maturity in the equine fetus. J. Royal Society Med.71:658-661.

152

Potter, G.D., J. W. Evans, G. W. Webb and S. P. Webb. 1987. Digestible energy requirements of Belgian and Percheron horses. Pages 133-138. In: Proc 10th Equine Nutr. Physiol. Symp, Ft. Collins, CO. Potter, G.D., S.P. Webb, J. W. Evans and G. W. Webb. 1989. Digestible energy requirements for work and maintenance of horses fed conventional and fat-supplemented diets. Pages 145-150. In: Proc. 11th Equine Nutr. Physiol. Sym., Stillwater, OK. Reed, K.R., and N.K. Dunn. 1977. Growth and development of the Arabian horse. Pages 76-90. In: Proc. 5th Equine Nutrition Symp. Reynolds, L.P., C. L. Ferrell, D.A. Robertson and S.P. Ford. 1986. Metabolism of the gravid uterus, foetus and utero-placenta at several stages of gestation in cows. J. Agric. Sci. 106:437-442. Scott, B.D., G.D. Potter, J.W. Evans, J.C. Reagor, G.W. Webb and S.P. Webb. 1989. Growth and feed utilization by yearling horses fed added dietary fat. J. Equine Vet. Sci. 9:210-214 Staniar, W.B., D.S. Kronfeld, K.H. Treiber, R.K. Splan and P.A. Harris. 2004. Growth rate consists of baseline and systematic deviation components in Thoroughbreds. J. Anim Sci. 82:1007-1015. Stillions M.C., and W.E. Nelson. 1972. Digestible energy during maintenance of the light horse. J. Anim. Sci. 34:981-982. Vermorel, M., and W. Martin-Rosset. 1997. Concepts, scientific bases, structure and validation of the French horse net energy system (UFC). Livestock Prod. Sci. 47:261-275 Vermorel, M., W. Martin-Rosset and J. Vernet. 1991. Energy utilization of two diets for maintenance by horses; agreement with the new French net energy system. J. Equine Vet. Sci. 11:33-35. Vermorel, M., W. Martin-Rosset and J. Vernet. 1997a. Energy utilisation of twelve forage or mixed diets for maintenance by sport horses. Livestock Prod. Sci 47:157-167. Vermorel, M., J. Vernet and W. Martin-Rosset. 1997b. Digestive and energy utilisation of two diets by ponies and horses. Livestock Prod. Sci. 51:13-19. Wooden G.R., K.L. Knox and C.L. Wild. 1970. Energy metabolism in light horses. J. Anim. Sci. 30:544548.

153

EQUINE CARBOHYDRATE NUTRITION: IMPLICATIONS FOR FEEDING MANAGEMENT AND DISEASE AVOIDANCE Raymond J. Geor, BVSc, PhD Middleburg Agricultural Research and Extension Center Virginia Polytechnic Institute and State University 5527 Sullivans Mill Rd Middleburg, VA 20117 Office: 540-687-3521 x26 Fax: 540-687-5362 Email: rgeor@vt.edu Summary This paper describes aspects of carbohydrate nutrition covered in the 6th revision of the Nutrient Requirements of Horses, including the classification and nomenclature of carbohydrates in horse feeds, a brief review of carbohydrate digestion in the horse, and the potential application of the glycemic index concept in equine nutrition. Perspectives on the implications for feeding management and disease avoidance are also presented. Introduction Carbohydrates are the primary source of energy in the diet of horses. As a non-ruminant herbivore, horses evolved to utilize forages high in structural carbohydrates, with bacterial fermentation and production of volatile fatty acids (VFA) in a highly developed large intestine. However, the modern horse, particularly those in athletic training, is often fed cereal grains or other starch-rich feeds to meet energy requirements. For example, survey studies have indicated that racehorses weighing 450 to 550 kg typically receive 3 to 6 kg of feed per day, with some horses receiving more than 8 kg per day (Gallagher et al., 1988; Southwood et al., 1993). Epidemiologic studies have reported that the feeding of grain or concentrates to horses may increase risk of colic (White 2006), and it has been hypothesized that disturbances to hindgut function associated with the feeding of starch-rich meals may be a contributing factor. There also is considerable speculation that diets high in starch and sugar might contribute to the pathogenesis of metabolic disorders, including osteochondrosis, insulin resistance and some forms of laminitis (Kronfeld and Harris, 2003). This hypothesis has stimulated research on the glycemic and insulinemic effects of different feeds, and driven commercial interest in the development of low starch feeds that, putatively, may offer health benefits to the horse. In recognition of the importance of carbohydrates in horse nutrition, the National Research Council (NRC) Subcommittee on Horse Nutrition included a chapter on carbohydrates in the 6th revision of the Nutrient Requirements of Horses. This paper presents some of the information on carbohydrate nutrition in that document, with emphasis on three areas: 1) the classification and nomenclature of carbohydrates in horse feeds, 2) aspects of carbohydrate digestion, and 3) application of the glycemic index concept in equine nutrition. Within each of these subsections, I offer some perspective on the possible implications for equine feeding programs. The material presented in these subsections is not contained within the 6th revision of the Nutrient Requirements of Horses and should not be taken as the opinion of the NRC Subcommittee on Horse Nutrition. Carbohydrates in Feeds Carbohydrates can be classified as mono-, di-, oligo- or polysaccharides. Monosaccharides of nutritional importance include glucose, fructose, and galactose. Free monosaccharides occur in low concentrations in plants, but they are important constituents of the oligosaccharides and polysaccharides
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

154

found in horse feeds. Two disaccharides are of nutritional importance: 1) lactose, a disaccharide composed of glucose and galactose, is a primary nutrient source for the nursing foal; and 2) maltose, a disaccharide consisting of two glucose units that is produced in the gastrointestinal tract by the action of amylase on starch. Oligosaccharides, which are short chains of monosaccharides, in animal feeds include raffinose, stachyose, and fructooligosaccharides (FOS), the latter also known as "fructans." Some fructans contain more than 10 monosaccharide units and are therefore classified as polysaccharides. The most common polysaccharides in horse feeds are starch and cellulose; pectin and hemicellulose are also polysaccharides. All carbohydrates contain similar amounts of gross energy. However, when utilized by the horse, they provide variable amounts of digestible energy, metabolizable energy or net energy. Carbohydrates digested and absorbed as monosaccharides in the small intestine yield more energy than carbohydrates digested by microbial action. The type of linkage between the monosaccharide residues in the oligosaccharides and polysaccharides affects the site of digestion of these compounds, and thus their nutritional value. Hydrolysis of the 1-6 and the 1-4 linkages of starch and maltose can occur in the equine small intestine, but horses do not produce the enzymes necessary to digest the 1-4 linkages found in cellulose or the mixed linkages found in hemicellulose. Therefore, digestion of cellulose and hemicellulose must occur as a result of microbial fermentation. Stachyose, raffinose, -glucans, FOS and pectin are also resistant to enzymatic hydrolysis. Several systems have been developed to classify carbohydrates. Because some carbohydrate components are difficult to measure, many carbohydrate partitioning systems have used collective terms such as nonstarch polysaccharides or total dietary fiber (Figure 1). The most common system of analysis for feeds is the system initially developed by Van Soest (1994). This system separates the feed into neutral detergent solubles and neutral detergent fiber (NDF). The NDF fraction contains cellulose, most hemicellulose, and lignin. Until recently, the amount of nonstructural carbohydrate (NSC) in a feed was determined by subtracting the amount of NDF, protein, ether extract, and ash from total dry matter. More recently, the term non-fiber carbohydrate (NFC) has been used to represent this difference, whereas NSC has been used to describe a chemically analyzed fraction of a feed (NRC, 2001). The NFC fraction is comprised of all carbohydrates not found in the NDF component of a feed. The NSC fraction includes mono- and disaccharides, oligosaccharides (including fructan) and starch (Hall, 2003). Few commercial feed analysis laboratories completely fractionate the carbohydrates that make up the NSC, but in most feeds the amount of NSC can be approximated by summing the amount of starch and the amount of water soluble carbohydrates (WSC). The extent to which the sum of starch and WSC accounts for all NSC depends on the analytical procedures used to measure these fractions. The quantitative difference between NSC and NFC is small for some feeds (e.g. cereal grains), but can be quite large for other feeds e.g. feeds with substantial pectin (Table 1). Table 1. Neutral detergent fiber (NDF), nonfiber carbohydrate (NFC), and nonstructural carbohydrate (NSC) composition of selected feedstuffs on a dry matter basis (NRC, 2001). Feedstuff % NDF % NFCb % NSCc Alfalfa Hay 43.1 22.0 12.5 Beet Pulp 47.3 36.2 19.5 Corn Gluten Meal 7.0 17.3 12.0 Mixed, Mostly Grass Hay 60.9 16.6 13.6 Soybean Meal (48% CP) 9.6 34.4 17.2 Soyhulls 66.6 14.1 5.3
The values shown here may vary from those shown elsewhere in this document. The values are provided to illustrate differences among feeds and carbohydrate categories. Actual values for individual feeds may vary by stage of maturity, variety, source, etc. b %NFC = (100% - %CP-%EE-%Ash-%NDF) c %NSC determined by measurement using the methods described by Smith (1981)
a

155

Figure 1. Fractionation of plant carbohydrates (from the 6th revision of the Nutrient Requirements of Horses).

Plant Carbohydrates1 Cell Contents


SugarsOligosaccharides3Fructan Glucans including mono & polysaccharides3 disaccharides fructooligosaccharides

Cell Wall
Pectins and Gums Hemicelluloses

Organic Acids2

Starch

Cellulose

Lignin2/Phenolics2

Nutritional/Digestive

Starch4

NSP TDF CHO-H4 NFC/NDSC6 CHO-FR5 CHO-FS5 NDF6 ADF CF8 NSP9 SDF10 TDF10

Starch4

WSC7 NSC7

Current and proposed systems for partitioning dietary carbohydrates based on current analytical methods (lower/shaded bracket; solid/dashed lines) and nutritional or physiologic definitions (upper bracket; dotted lines) relative to equine digestive function. Adapted from Hall, 2003 and Hoffman et al., 2001. In the analytical bracket, dashed lines indicate that recovery of included compounds may be incomplete. Abbreviations: ADF =acid detergent fiber; CF=crude fiber; CHO-H=hydrolyzable carbohydrates; CHO-Fs=slowly fermented carbohydrates; CHO-FR = rapidly fermented carbohydrates; NDF=neutral detergent fiber; NDSC=neutral detergent soluble carbohydrates; NFC= non-fiber carbohydrates; NSC= non-structural carbohydrates; NSP=nonstarch polysaccharides; SDF=soluble dietary fiber; TDF= total dietary fiber; WSC= water-soluble carbohydrates 1. Major categories of carbohydrates and associated substances are shown. These categories may not include all carbohydrates produced by plants. 2. Some non-carbohydrate components are included here as they are components of the specific analytical fractions. 3. Specific fructans can be categorized as either fructooligosaccharides or fructan polysaccharides depending on degree of polymerization. 4. A variable fraction of total starch can be resistant to enzymatic hydrolysis and thus some starch may appear in other nutritional fractions. 5. Fermentability of gums may be variable. 6. Some hemicellulose may be soluble in neutral detergent and thus recovered in the NFC/NDSC fraction, rather than the NDF fraction. 7. Recovery of compounds in the analytical WSC fraction (and thus the NSC fraction when NSC is approximated as starch + WSC) may depend on methodology used. 8. Amount of cell wall constituents included in CF analysis varies by feed. 9. From a nutritional perspective, NSP includes all polysaccharides except starch. However, the analytical method for NSP may recover a variable amount of fructan polysaccharide. 10. From a nutritional perspective, TDF includes all carbohydrates resistant to mammalian digestion. However, the analytical method for TDF (and SDF) does not recover oligosaccharides and may recover a variable amount of fructan polysaccharides.

Analytical

156

The system of partitioning carbohydrates into NDF and NFC fractions was developed for use in ruminant nutrition. Because NDF and NFC are heterogeneous mixtures of carbohydrates that vary in digestibility, additional partitioning of carbohydrate fractions may be useful. Hoffman et al. (2001) proposed that a relevant system for partitioning carbohydrates in equine feeds would include three main fractions: (1) hydrolyzable carbohydrates (CHO-H), which can be digested in the small intestine; (2) rapidly fermented carbohydrates (CHO-FR), which are readily available for microbial fermentation; and (3) slowly fermentable carbohydrates (CHO-FS). These authors suggested the hydrolyzable fraction included hexoses, disaccharides, some oligosaccharides, and the nonresistant starches. Although some fermentation of these compounds may occur in the stomach, the primary products of digestion of these compounds are monosaccharides, and thus the energy yield is relatively high. The rapidly fermentable fraction included pectin, fructan, and some oligosaccharides not digested in the small intestine. Resistant starch and neutral detergent hemicellulose could also be included in the rapidly fermented fraction. The slowly fermented carbohydrate fraction includes cellulose, hemicellulose, and ligno-cellulose that result primarily in the production of acetate in the large intestine. The system proposed by Hoffman et al. (2001) may allow better understanding of the energy value of the carbohydrate portion of the feed as it separates the components that are absorbed as glucose from those that are fermented to VFA. Unfortunately, methods for analyzing all of these fractions are not readily available. Figure 1 shows a schematic of the various analytical and nutritional methods for classifying carbohydrates. Perspectives Improved understanding of the carbohydrate composition and nutritional value of equine forages and feeds should facilitate a more accurate estimation of energy content and, with more detailed information on the digestion and metabolism of the carbohydrate fractions, may be useful in the development of feeding recommendations that lessen risk of digestive and metabolic disorders. In this context, the carbohydrate partitioning scheme proposed by Hoffman and colleagues (2001) may offer advantages over systems based on standard proximate analysis, but practical application will be limited unless methods for component analysis of the carbohydrate fractions become more widely (commercially) available. Further work is also needed to determine the digestive fate of different carbohydrates in equids, in particular better characterization of the rapidly fermentable (CHO-FR) and slowly fermentable (CHOFS) fractions as described by Hoffman et al. (2001). The detailed description of carbohydrate nomenclature (see Fig. 1) in the new NRC document may alleviate some of the existing confusion in this area. Currently, some commercial laboratories measure WSC (i.e. free sugars plus fructans) but report it as sugar, while others use the term sugar to describe the free sugar fraction only. A more accurate and consistent description of the measured (or calculated) fractions is needed. Perhaps also commercial forage laboratories will adopt other enzymatic or chromatographic techniques for measurement of fructans, which have been implicated in the pathogenesis of pasture-associated laminitis. Carbohydrate Digestion Structural carbohydrates are important sources of energy in horse diets. The microbial production of VFA in the cecum may be sufficient to meet up to 30 percent of a horses energy needs at maintenance (Glinsky et al., 1976), with additional VFA are produced in the colon. Therefore, the contribution of VFA to total daily energy utilization is substantial, particularly for horses consuming all-forage diets. Acetate is the principal VFA produced, but propionate and butyrate production may also be significant (Argenzio and Hintz, 1970; Hintz, et al., 1971; Moore-Colyer et al., 2000). Most of the energy in grains is found as starch. Total tract starch disappearance in the horse is very high (> 90%), but the extent of prececal starch digestion is variable. There is considerable structural variation in the starch found in different grains (Kienzle et al., 1997). The two main components of starch are amylose and amylopectin. Amylose is a straight chain structure of glucose units, while amylopectin is a branched chain of glucose units. The proportions of amylose and amylopectin in starch vary by cereal grain and with other factors, including maturity and variety (Van Soest, 1994). The starch that reaches the 157

large intestine is categorized as resistant starch in human nutrition (Englyst and Englyst, 2005). Resistant starch includes starch that is not accessible to digestive enzymes due to its structure or encapsulation in plant cell structures and starch that has been modified by certain types of processing (Cummings et al., 1996; Tharanathan, 2002). A system for characterizing the availability of the starch found in various horse feeds has not been developed, but could be a useful tool. Enzymatic digestion of starch occurs in the small intestine, but starch is also susceptible to microbial fermentation. Healy et al. (1995) reported increased lactate concentrations in gastric fluid obtained from ponies fed a large concentrate meal and other researchers have reported significant dry matter and starch disappearance to occur in the stomach (de Fombelle et al., 2003; Varloud et al., 2003). It is also possible that fructan can be hydrolyzed to some extent by gastric acid (Van Soest, 1994) or fermented in the stomach. The type and amount of forage in the diet may alter small intestinal starch digestibility (Meyer et al., 1993), possibly by affecting digesta flow in the small intestine. The abundance of amylase in the digestive secretions may also influence starch digestibility in the small intestine. However, little is known about the factors that affect the quantity of digestive enzymes secreted into the equine gut. The amount of amylase measured in the pancreatic tissue of horses fed either hay or hay and concentrate for at least 8 weeks was not affected by diet (Kienzle et al., 1994). However, in another experiment, the amylase activity of jejunal chyme was higher when horses received a diet with added corn, oats, or barley compared to only hay (Kienzle, et al., 1994). There are differences in prececal starch disappearance among the different grains (Radicke et al., 1991; Kienzle et al., 1992; de Fombelle et al., 2004), with oat starch generally being more digestible in the small intestine than corn starch or barley starch. The amount of starch consumed at one time may also affect the percentage of starch that disappears before reaching the large intestine. When a small amount of oats was fed to horses, approximately 80% of the starch was digested and absorbed prior to reaching the terminal ileum, but when a larger amount of oats was fed, starch disappearance before reaching the ileum was only 58% (Potter et al., 1992). Starch not digested in the small intestine enters the large intestine where it will be fermented, yielding less net energy than when it is absorbed as glucose. Starch flow to the large intestine may alter the microbial environment and have adverse effects on digestive function. Feeding a high-starch diet (30% starch; 3.4 g starch/kg BW/meal) has been reported to decrease the concentration of celluloytic bacteria in the cecum (Medina et al., 2002). In addition, several studies have reported decreased cecal and/or colonic pH in horses or ponies fed high starch diets (Willard et al., 1977; Radicke et al., 1991; Medina et al., 2002). To avoid a decrease in large intestinal pH, high-starch feeds should not be offered in amounts that result in substantial starch overflow to the large intestine. One researcher has suggested that the capacity of the small intestine for starch digestion is reached at a starch intake of 3.5 to 4.0 g/kg BW (Potter et al., 1992). However, Radicke et al. (1991) found that cecal pH was suppressed with starch intakes between 2 and 3 g/kg BW. Perspectives The overfeeding of starch (grain overload) has been put forth as a risk factor for colic in horses, presumably related to starch overflow to the large intestine. The daily feeding of concentrate at 2.5 to 5.0 kg/day and > 5.0 kg/day to mature horses increased the risk of colic 4.8 and 6.3 times, respectively, when compared to horses fed no concentrate (Tinker et al., 1997). In other studies, the feeding of more than 2.7 kg oats per day or a recent change in grain feeding (within the last 2 weeks) was associated with increased risk of colic (White, 2006). These data justify recommendations for careful dietary management of horses receiving grain-concentrate that might lower risk of digestive disturbance and colic, e.g. feeding starch sources with high prececal digestibility, limiting the size of individual starch-rich meals (no more than 2-3 g starch/kg bwt.), use of alternative sources of energy (e.g. vegetable oils) with a decrease in starch feeding, and/or use of dietary additives such as yeast cultures that are purported to mitigate decreases in hindgut pH associated with starch overload. Anecdotally, these measures have been helpful in decreasing the incidence of colic on individual farms, but well controlled, prospective studies are needed for rigorous evaluation of these recommendations.

158

Glycemic and Insulinemic Responses Application of a Glycemic Index to Horse Feeds There is interest in the effects of diet on postprandial glucose and insulin responses in horses. In human nutrition, the term glycemic index (GI) has been used to characterize the magnitude of the blood glucose response after consumption of various foods. The GI has been defined as the incremental area under the blood glucose response curve of a 50 g carbohydrate portion of a test food, expressed as a percentage of the blood glucose response curve to the same amount of carbohydrate from a standard food consumed by the same person (FAO/WHO, 1998). The 50 g carbohydrate portion should contain 50 g of available carbohydrate, and the standard food may be either white bread or glucose (FAO/WHO, 1998). The FAO/WHO guideline suggested testing the standard food at least three times in each subject to obtain a representative mean response. However, it should be recognized that the GI determined with a specific amount of carbohydrate may not represent responses to other levels of intake. For this reason, the term glycemic load (GL) was introduced to quantify the overall glycemic effect of a meal, which is a function of both the type and quantity of carbohydrate in the food or meal. More specifically, the GL is defined as the product of a foods GI and its total available carbohydrate content. Several studies have attempted to apply the GI concept to horse feeds. However, interpretation and comparison of the results are hampered by the wide variation in methodology. The glucose response to the carbohydrate in test feeds has been compared to the response to an equivalent amount of glucose administered by gastric gavage (Jose-Cunilleras et al., 2004) or to a meal of corn or oats (Pagan et al., 1999; Vervuert et al., 2003; Vervuert et al., 2004; Rodiek and Stull, 2005; Vervuert et al., 2005). A major source of variation among studies has been the amount of feed or carbohydrate offered. Among studies that have estimated the GI of feeds, there has been little consistency in the amount of carbohydrate offered to horses. In one study, the amount of hydrolyzable carbohydrate was approximately 900 g/484 kg horse (Jose-Cunilleras et al., 2004). In other studies, smaller amounts of hydrolyzable carbohydrates have been fed (Vervuert et al., 2003, 2004). When the glycemic responses to several feeds were compared to oats, the relative GI values were affected by the amount of the test feeds and the oats that were offered (0.75, 1.5, or 2.5 kg) (Pagan et al., 1999). Feeds have not always been offered in equivalent carbohydrate amounts or meal sizes. One study compared feeds that provided similar amounts of calories that resulted in large differences in both total hydrolyzable carbohydrate intake and total feed intake among feeds (Rodiek and Stull, 2005). An advantage of this method would be that feed substitutions in horse diets are often made on a caloric basis. However, a disadvantage relates to the time required to consume the feed provided and the effects this may have on glycemic response. Perspectives Do the GI and GL concepts have a place in equine nutrition? In human nutrition, the GI system has been used in formulating dietary recommendations for diabetic patients for which moderation of postprandial hyperglycemia is desirable. Similarly, it has been proposed that the feeding of high glycemic diets contributes to the pathogenesis or clinical manifestations of several endocrine and metabolic conditions in equids (e.g. some forms of laminitis, pituitary dysfunction), justifying the formulation of diets that produce attenuated glycemic and insulinemic responses. In reality, the evidence linking high glycemic diets with disease is very limited, although the clinical expression of some forms of exertional rhabdomyolysis, notably polysaccharide storage myopathy (PSSM) in Quarter horses, is related to the amount of dietary NSC. Appropriate formulation of diets for affected animals requires knowledge of the amount of NSC in the feed and the glycemic/insulinemic response following consumption of that feed. In this context, there is a place for the GI and GL concepts in equine dietetics but a few hurdles must first be overcome. A suitable assay for the determination of available carbohydrate in horse feeds is needed. The NSC content of a feed has been used for this purpose but this fraction can include carbohydrates (e.g. fructans) not subject to enzymatic digestion in the small intestine. Furthermore, studies in our laboratory have indicated that NSC content is a poor predictor of meal glycemic response. Second, there is a need for more exploration of physiological factors underlying variation in glycemic and insulinemic responses. Studies to date have indicated huge variation in the glycemic response to similar feeds. Differences in research protocols and the chemical composition of the feeds may account for some 159

of this variation, but animal factors (e.g. rate of gastric emptying, insulin sensitivity) also may contribute. Development of a standardized methodology for assessment of glycemic responses would be helpful for studies that aim to characterize the sources of variation in glycemic response, and is critical to development of a valid GI system. It also should be recognized that the glycemic response to a mixed meal is not easily predicted from the GIs of the individual feed components. Thus, a GI system for horses also should account for responses of horses consuming mixed meals. References Argenzio, R., and H.F. Hintz. 1970. Glucose tolerance and effect of volatile fatty acid on plasma glucose concentration in ponies. J. Anim. Sci. 30:514-519. Cummings, J.H., E.R. Beatty, S.M. Kingman, S.A. Bingham, and H.N. Englyst. 1996. Digestion and physiological properties of resistant starch in the human large bowel. Br. J. Nutr. 75:733-747 de Fombelle, A., A.G. Goachet, M. Varloud, P. Boisot and V. Julliand. 2003. Effects of diet on prececal digestion of different starches in the horse measured with the nylon bag technique. Pages 115116. In: Proc. 18th Equine Nutr. Physiol. Soc. Symp., East Lansing, MI. de Fombelle, A., L. Veiga, C. Drogoul and V. Julliand. 2004. Effect of diet composition and feeding pattern on the prececal digestibility of starches from diverse botanical origins measured with the mobile nylon bag technique. J. Anim. Sci. 82:3625-3634. Englyst, K.N., and H.N. Englyst. 2005. Carbohydrate availability. Br. J. Nutr. 94:1-11. FAO/WHO (Food and Agriculture Organization/World Health Organization). 1998. Carbohydrates in human nutrition (FAO Food and Nutrition Paper 66). Available at http://www.fao.org/docrep/W8079E/w8079e00.htm Glinsky, M.J., R.M. Smith, H.R. Spires, and C.L. Davis. 1976. Measurement of volatile fatty acid production rates in the cecum of the pony. J. Anim. Sci. 42:1465-1470. Hall, M.B. 2003. Challenges with nonfiber carbohydrate methods. J. Anim. Sci. 81:3226. Healy, H.P., P.D. Siciliano and L.M. Lawrence. 1995. Effect of concentrate form on blood and gastric fluid variables in ponies. J. Equine Vet. Sci. 15:423-428. Hintz, H.F., R.A. Argenzio, and H.F. Schryver. 1971. Digestion coefficients, blood glucose levels, and molar percentage of volatile fatty acids in intestinal fluid of ponies fed varying forage-grain ratios. J. Anim. Sci. 33:992-995. Hoffman, R.M., J.A. Wilson, D.S. Kronfeld, W.L. Cooper, L.A. Lawrence, D. Sklan, and P.A. Harris. 2001. Hydrolyzable carbohydrates in pasture, hay and horse feeds: direct assay and seasonal variation. J. Anim. Sci. 79:500-506. Jose-Cunilleras, E., L.E. Taylor, and K.W. Hinchcliff. 2004. Glycemic index of cracked corn, oat groats and rolled barley in horses. J. Anim. Sci. 82:2623-2629. Kienzle, E., J. Pohlenz, and S. Radicke. 1997. Morphology of starch digestion in the horse. J. Vet. Med. A 44:207-221. Kienzle, E., S. Radicke, W. Wilke, E. Landes, and H. Meyer. 1992. Preilieal starch digestion in relation to source and preparation of starch. Pferdeheilkunde (1. European Conference on Horse Nutrition):103106. Kienzle, E., S. Radicke, E. Landes, D. Kleffken, M. Illenseer and H. Meyer. 1994. Activity of amylase in the gastrointestinal tract of the horse. J. Anim. Physiol A. Anim Nutr. 72:234-241. Kronfeld, D.S., and P.A. Harris. 2003. Equine grain-associated disorders. Comp. Cont. Educ. Pract. Vet. 25:974-982. Medina, B., I.D. Girard, E. Jacotot, and V. Julliand. 2002. Effect of a preparation of Sacchromyces cervisiae on microbial profiles and fermentation patterns in the large intestine of horses fed a high fiber or a high starch diet. J. Anim. Sci. 80:2600-2609. Mertens, D.R. 1997. Creating a system for meeting the fiber requirements of dairy cows. J. Dairy Sci. 80:1463-1481. Moore-Colyer, M.J.S., J.J. Hyslop, A.C. Longland, and D. Cuddeford. 2000. Intra-cecal fermentation parameters in ponies fed botanically diverse fibre-based diets. Anim. Feed Sci. Tech. 84:183-197.

160

NRC (National Research Council). 2001. Nutrient Requirements of Dairy Cattle. Washington, DC: National Academy Press. NRC (National Research Council). 2007. Nutrient Requirements of Horses, 6th Revision. Washington, DC: National Academy Press. Pagan, J.D., P.A. Harris, M.A.P. Kennedy, N. Davidson, and K.E. Hoekstra. 1999. Feed type and intake affect glycemic response in Thoroughbred horses. Pages147-149. In: Proc. 1999 Equine Nutr. Conf. Feed Manufact., Lexington, KY. Potter, G., F. Arnold, D. Householder, D. Hansen, and K. Bowen. 1992. Digestion of starch in the small or large intestine of the equine. Pferdeheilkunde. (1. European Conference on Horse Nutrition):109111. Radicke, S., E. Kienzle, and H. Meyer. 1991. Preileal apparent digestibility of oats and cornstarch and consequences for cecal metabolism. Pages 43-48. In: Proc. 12th Equine Nutr. Physiol. Soc. Symp., Calgary, Alberta. Rodiek A.V., and C. Stull. 2005. Glycemic index of common horse feeds. Pages 154-155. In: Proc. 19th Equine Science Symp., Tucson, AZ. Southwood, L., D.L. Evans, W.L. Bryden, and R.J. Rose. 1993. Nutrient intake of horses in Thoroughbred and Standardbred stables. Aust. Vet. J. 70:164-168. Tinker, M.K., N.A. White, P. Lessard, C.D. Thatcher, K.D. Pelzer, B. Davies, and D.K. Carmel. 1997. Retrospective study of equine colic risk factors. Equine Vet. J. 29:454-458. Van Soest, P.J. 1994. Carbohydrates. Pages 156-176. In: Nutritional ecology of the ruminant, 2nd Ed Cornell University Press, Ithaca, NY. Varloud, M., A.G. Goachet, A. de Fombelle, A. Guyonvarch and V. Julliand. 2003. Effect of the diet on prececal digestibility of the dietary starch measured in horses with acid insoluble ash as an internal marker. Pages 117-118. In: Proc. 18th Equine Nutr. Physiol. Soc. Symp., East Lansing MI. Vervuert, I., M. Coenen, and C. Bothe. 2003. Effects of oat processing on the glycaemic and insulin responses in horses. J. Anim. Physiol. Anim Nutr. 87:96-104. Vervuert, I., M. Coenen, and C. Bothe. 2004. Effects of corn processing on the glycaemic and insulinaemic responses in horses. J. Anim. Physiol. Anim. Nutr. 88:348-355. Vervuert, I., M. Coenen, and C. Bothe. 2005. Glycaemic and insulinaemic indexes of different mechanical and thermal processes grains for horses. Pages 154-155. In: Proc. 19th Equine Science Symp., Tucson, AZ. White, N.A. 2006. Equine colic. II. Causes and risk factors for colic. Proc. 52nd Amer. Assoc. Eq. Pract. 52:115-119. Willard, J. G., J.C. Willard, S. A. Wolfram and J. P. Baker. 1977. Effect of diet on cecal pH and feeding behavior of horses. J. Anim. Sci. 45:87-93.

161

PROTEIN AND AMINO ACID NUTRITION IN THE HORSE: IMPROVEMENTS AND GOALS FOR THE FUTURE Patty Graham-Thiers Virginia Intermont College Campus Box 317 1013 Moore St. Bristol, VA 24201 Phone: 276-466-7168 FAX: 276-669-5763 Email: thiers@vic.edu Summary Protein is an important nutrient for tissue synthesis in the body. In previous editions of the NRC, protein requirements for horses have been based mostly on a crude protein:digestible energy ratio. In the current edition of the NRC (2007), requirements have been made based on replacing protein losses or providing additional protein for the various functions where the data were available to calculate these needs. Research is very limited regarding specific amino acid requirements for the horse. Variations in digestibility have no doubt clouded the ability to determine protein requirements across all potential sources of protein. The differences in amino acid profiles among protein sources also makes expressing protein requirements on a crude protein basis seem an oversimplification especially when the horses actual need is for amino acids, not just protein. These differences in digestibility as well as varying amino acid profiles of various protein sources make additional research a necessity. This is true not only of the needs of the horse but for advances in determining protein that is available from various feedstuffs. If we are going to progress to expressing our protein requirements on a digestible protein basis rather than a crude protein basis these areas need to be explored. Introduction Protein is used to build and repair tissues, synthesize hormones and enzymes. The horse breaks down protein in the foregut (stomach and small intestine) into amino acids that are then absorbed into the bloodstream and become part of the amino acid pool. Protein entering the hindgut is processed by microbes that produce ammonia and microbial protein. The available evidence does not find a significant contribution of amino acids to the amino acid pool from hindgut protein digestion in the horse. Therefore, the source of protein provided in the horses diet should be considered for its amino acid profile and digestibility in the foregut. Protein recommendations in the 1989 version of the National Research Council (NRC) estimated protein requirements based on a crude protein (CP):digestible energy (DE) ratio calculated with data from various studies that were available. This resulted in a CP:DE ratio for maintenance and exercise of 40g CP/Mcal DE and thus requirements were based on this relationship. The ratio was increased slightly for gestation to 44g CP/Mcal DE during the last 3 months of gestation based on the assumption that additional protein would be needed for fetal development. Lactation CP recommendations were based on milk production and milk protein concentrations to estimate the amount of protein necessary for normal milk production. Growth data that were available were used to estimate an ideal CP:DE ratio of 50 g CP/Mcal DE for weanlings and 45 g CP/Mcal DE for yearlings based on maximizing average daily gain (ADG). Requirements for lysine were based on the assumption that diets that provided adequate amounts of CP generally contained lysine concentrations of 3.5% of the CP. In the case of growth, data that were available observed maximal ADG at a calculated lysine:DE ratio of 2.1 g lysine/Mcal DE for weanlings and 1.9 g lysine/Mcal DE for yearlings. No other official recommendations were made for lysine
Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

162

requirements. All other functions such as pregnancy, lactation and exercise based the requirement for lysine on 3.5% of the CP being lysine. In the current version of the NRC (2007), the CP requirements were based on actual protein needed for a particular function, overall digestibility and efficiency of use of the digestible protein rather than a CP:DE ratio. Linear and non-linear regression was applied to the data from available studies to estimate protein losses, digestibility and efficiency of use to make the current CP and lysine recommendations. In instances when adequate data were available, lysine requirements were determined via linear regression. For all other functions, the lysine requirement was estimated as a percentage of the CP requirement. Crude Protein and Lysine Requirements Maintenance Protein is 16% nitrogen (N). Nitrogen balance is often used to evaluate N loss, N retention and therefore CP requirements. Nitrogen balance studies have concluded that between 400 and 800 mg digestible protein (DP)/kg BW/day is necessary to achieve nitrogen balance in the sedentary horse (Slade et al., 1970; Hintz and Schryver, 1972; Harper and Vander Noot, 1974; Reitnour and Salsbury, 1976; Meyer, 1985; Patterson et al., 1985; Olsman et al., 2003). The NRC (1989) CP requirement for maintenance (656 g CP/day for 500-kg horse) assumed a 46 percent digestibility and an all-forage diet. This equates to approximately 600 mg DP/kg BW/day. Applying linear regression to means from studies that measured nitrogen intake and nitrogen retention resulted in 619 mg DP/kg BW/day (813 mg CP/kg BW/day) to be necessary for zero nitrogen retention (r2 = 0.76). Thus, based on these data, the minimum digestible protein intake for horses in maintenance should be > 620 mg DP/kg BW/day. Fitting the same data to a two-slope model estimates the requirement to be 0.202 g N/kg BW/day which would result in some nitrogen reserves in the body rather than a zero balance. The CP equivalent for this amount of N is 1.26 g/kg BW/day. This equals 630 g CP for the 500-kg horse (Table 1). Lysine requirements for horses in maintenance have not been specifically studied however, using means from studies that reported diet composition as well as intake and N retention, an intake of 0.036 g lysine/kg BW/day was calculated to result in zero N retention. The two-slope model identified a plateau in N retention for maintenance horses at an intake of 0.054 g lysine/kg BW/day. This equates to a minimum of 18 g lysine per day and an optimum of 27 g lysine per day for the 500-kg horse. This is also equivalent to 4.3% of the CP being lysine (27g lysine/630g CP). The quality of the protein source needs to be considered in order to supply adequate lysine. Growth Protein in the growing horse would be used primarily for tissue formation and be reflected in the horses average daily gain (ADG). By combining data from several studies that evaluated growth in weanling horses, an estimated optimum intake of 4 g CP/kg BW/day for weanlings between 4 and 10 months of age with an expected mature body weight of 500 kg was observed. This equals 672 g CP per day for the 4-month-old (168-kg) weanling. After accounting for the CP need for maintenance, a 50 percent efficiency of use is estimated from the remaining CP for gain (0.84 kg/day) assuming the gain is 20 percent protein. Allowing for an adjustment for digestibility, CP requirements can be estimated from body weight, rate of gain and efficiency of use of the digested CP. A growth curve was estimated from available growth data to estimate body weight and ADG for growing horses of various ages (Table 1).

163

Table 1. Comparison of daily CP and lysine requirements from the 5th and 6th edition of the Nutrient Requirements of the Horse. CP, g (1989)1 Maintenance minimum average elevated 656 CP, g (2007)2 540 630 720 630 685 704 729 759 797 841 893 lysine, g (1989)1 23 lysine, g (2007)2 23 27 31 27 29 30 31 33 34 36 38

Pregnant Early 5m 6m 7m 8m 9m 10 m 11 m Lactation First 3 months Last 3 months 1st month 2nd month 3rd month 4th month 5th month 6th month Working Light Moderate Intense(heavy) Very Heavy Weanling 4 months 6 m, moderate growth 6 m, rapid growth Yearling 12 months 12 m, in training 18 months
1 2

801 815 866 1427 1048

28 29 30 50 37

1535 1530 1468 1398 1330 1265 820 984 1312 699 768 862 1004 669 676 29 34 46

85 84 80 76 71 67 30 33 37 43 29 29

720 750 860 851 956 893

30 32 36 36 40 38

846 799

36 34

NRC (1989) NRC (2007)

De Almeida et al. (1998) observed that nitrogen retention was maximized at a protein intake of 3.2 g CP/kg BW/day (2.4 g DP/kg BW/day) for yearling horses. Yearlings (315-333 days of age) have had greater ADGs when fed at least 3.3 g CP/kg BW/day when fed soybean meal (SBM) and alfalfa hay (Ott and Kivipelto, 2002), and 3 g CP/kg BW/day when fed SBM and bermudagrass hay (Ott and Asquith, 1986). Analysis of means from studies providing intake data reported improved ADG with increasing CP intake up to 3.3 g CP/kg BW/day for yearlings between 11 and 17 months of age. Calculating the efficiency of use of CP for gain in studies reporting CP intake, BW and ADG resulted in average efficiency of only 30 percent for horses over 11 months of age. A 321-kg yearling (12 months

164

old) gaining 0.45 kg/day would require 462 g CP/day for maintenance needs and 380 g CP per day for gain after adjusting for efficiency of use and digestibility assuming gain is 20 percent protein. This would result in a total CP recommendation of 842 g CP/day for the yearling with an expected mature weight of 500 kg. Studies evaluating lysine needs of growing horses have concluded that lysine intake for weanlings (4 to 10 months of age) should be 33 to 42 g/day (151 to 179 mg lysine/kg BW/day) to improve ADG (Ott et al., 1979, Ott and Kivipelto, 2002; Breuer and Golden, 1971). Using the means from studies that reported ADG as well as diet, body weight, and feed intake, estimates of amino acid intake can be made. The two-slope model estimated the lysine requirement to be 168 mg/kg BW/day. For the 4-month-old 168-kg weanling, this would be 28 g lysine per day. This amount of lysine is equivalent to 4.3 percent of the horse's CP requirement. Therefore, the lysine requirement for weanling horses between 4 and 10 months of age is 4.3 percent of the CP requirement. This recommendation is also used for yearling horses as well (Table 1). Pregnancy Protein and amino acid requirements for the pregnant mare have received very little attention in the literature. The lower amount of acceptable protein intake for mares in early to mid gestation has not been investigated thoroughly. In a study by van Niekerk and van Niekerk (1997) the CP:DE ratio was <35 in mares that lost weight and had higher incidence of fetal loss compared to the other groups with CP:DE ratios 38. A study by Boyer et al. (1999) had a CP:DE ratio of 40 and observed all mares in positive nitrogen balance. It can be concluded that feeding mares in early- to mid-gestation at an average maintenance level of CP intake is adequate however, more research needs to be done to determine more precisely the needs of the mare in early- to mid-pregnancy. The current NRC (2007) recommendations have a CP: DE ratio of approximately 42 g CP/Mcal DE. Fetal growth rate has been estimated in the mare (Platt, 1984). Foal body composition has also been evaluated (Doreau et al., 1986). Based on estimated protein composition of the fetus (20%), as well as the rate of fetal gain, estimates of protein needs above maintenance were made with an allowance for placental and uterine protein needs included. This attempted to account for additional protein needs for gestation rather than simply assuming that increased caloric intake would provide adequate CP for the mare during pregnancy. Without any available data regarding the amino acid needs of the pregnant mare, the lysine requirement for pregnancy was estimated to be 4.3 percent of the CP requirement (Table 1). Lactation The amount of milk production by mares has been documented in several studies and reported to vary between 1.9 and 3.9 percent of the mares body weight throughout lactation (up to six months). The same studies have reported protein content of milk: from 3.1 to 3.3 percent in early lactation (generally colostrums) and gradually declining to 1.6 to 1.9 percent in later lactation (Doreau et al., 1986; 1990; Smolders et al., 1990; Martin et al., 1992; Mariani et al., 2001). Therefore, based on milk production, milk protein concentrations, efficiency of use and digestibility an additional 50g CP was determined to be needed per kg of milk produced above maintenance (Table 1). Wickens et al. (2002) determined the amino acid profile in mares milk and formulated an ideal protein (amino acid requirements) from that data. Assuming an average lysine content of 1.7 g/kg milk and 65 percent utilization efficiency, a dietary lysine requirement was estimated to be 2.62 g digestible lysine/kg milk. Adjusting for digestibility, 3.3 g lysine would be needed per kg milk produced over maintenance needs for lysine (Table 1). Both estimates for CP and lysine requirements for the lactating mare result in dietary quantities that may be difficult to provide in the short term. It is presumed that the lactating mare readily utilizes

165

body reserves to overcome short-term deficiencies. Careful selection of protein sources during this demanding time is certainly warranted. Exercise There is some evidence that the exercising horse requires additional protein per kilogram of body weight for developing muscle and repair of damaged muscle. Freeman et al. (1988) reported an increase in nitrogen retention for exercising horses as exercise load increased. Horses retained an additional 0.37 g CP/kg BW/day during exercise compared to rest periods. However, this study did not account for nitrogen lost in sweat. Estimating the amount of protein lost in sweat in this study could account for 0.23 g of CP/kg BW out of the additional 0.37 g CP/kg that was apparently retained. This leaves 0.14 g CP/kg BW unaccounted for in nitrogen losses and may represent the additional protein need over maintenance for exercising horses. Using this data would result in a CP recommendation of 1,000 g CP for the 500 kg horse in similar work as was observed in this study (very heavy intensity). Wickens et al. (2003) observed exercising horses (heavy intensity) to have maximal nitrogen retention when horses were fed a diet providing 1,016 g CP/day. Wickens et al. (2005) also utilized 3methylhistidine (3MH) to estimate protein requirements for exercising horses. Muscle protein turnover can be estimated using 3MH and these data predicted a crude protein requirement of 954 g/d. These data and those from the studies of Freeman et al. (1988) result in a recommendation of 1.9 to 2.1 g CP/kg BW/day for the heavily exercised horse. For the 500-kg horse this would result in a requirement of 950 to 1,050 g CP/day. The protein requirement for the exercising horse is therefore based on the fact that additional muscle appears to be gained during conditioning and that nitrogen is lost in sweat. Both muscle tissue gain and sweat loss are dependent on the intensity of exercise. Adjustments were also made for efficiency of CP use and digestibility (Table 1). Using means from studies that reported diet and intake, lysine intake was calculated. A two-slope model was used to analyze means of lysine intake and nitrogen retention in exercising horses. The analysis resulted in a recommendation of 0.068 g lysine/kg BW/day for the exercising horse. This would result in a recommended lysine intake of 37 g/day for the 500-kg horse. The horses in these studies participated in high intensity exercise. With a CP requirement of 862 g CP/day for the 500-kg horse in heavy exercise, the lysine requirement of 37 g/day represents 4.3 percent of the CP requirement. Therefore, the requirement for lysine for the exercising horse is based on 4.3% of the CP requirement (Table 1). The Future Digestibility Crude protein is calculated by multiplying the concentration of nitrogen by 6.25 since protein is 16% nitrogen. It would appear that one can presume that any source of protein that can provide adequate CP to the horse would be acceptable. The horse in maintenance requires 630 g CP. Based on the previous assumption, this quantity of CP could be provided by 8.2 kg of grass hay at 8% CP, 1.4 kg of soybean meal or 1.5 kg sunflower meal. The problem with this is that it does not take into account the difference in digestibility nor does it consider the amino acid profile of the protein source. The same quantities of feedstuffs needed to provide 630 g CP would provide 32.8 g, 39.2 g and 15 g lysine respectively for grass hay, soybean meal and sunflower meal. Despite soybean meal and sunflower meal have very similar concentrations of CP, they provide very different quantities of lysine due to differences in their amino acid profiles (Table 2). By continuing with CP as the unit of measure for protein requirements for the horse, we are ignoring variances in protein chain structure and the availability (via digestion) of the source of protein to the horse.

166

Table 2. Comparison of CP and lysine concentrations in various feedstuffs.1 Feedstuff Soybean meal Sunflower meal Grass Hay 1 NRC, 2007 CP, % 44 42 8 Lysine % 2.8 1.0 0.4 CP, g (1 kg) 440 420 80 Lysine, g (1 kg) 28 10 4

Total tract apparent digestibility of CP varies based on protein source (e.g., fishmeal vs. corn gluten meal) and components of the diet (e.g., forage vs. concentrate) as well as with the ratio of forage to concentrate in the daily ration. Studies evaluating the CP digestibility of forages found apparent total tract CP digestibilities varying from 73 to 83 percent for alfalfa hay, 57 to 64 percent for Coastal Bermudagrass, and 67 to 74 percent for other grasses such as fescue and bromegrass (Gibbs et al., 1988). Digestibility of concentrates such as corn, oats, or sorghum were found to have apparent total tract protein digestibilities of 88.0, 82.8, and 84.6 percent, respectively when fed with grass hay (Farley et al., 1995). A similar trial was conducted to evaluate protein digestibility of protein supplements such as soybean meal (SBM) and cottonseed meal (CSM). Calculating the apparent digestibility of the protein supplement by difference resulted in total tract digestibility of 92.2 and 85.0 percent for SBM and CSM, respectively (Potter et al., 1992). Compiling means from studies that have reported nitrogen intake as well as fecal nitrogen, resulted in an estimate of apparent nitrogen digestibility of 79 percent for the total tract when linear regression was applied to the data (r2 = 0.94) Digestible Protein/Available Protein Understanding protein digestibility in the horse enables protein requirements to be expressed in terms of digestible protein (DP) instead of CP. When protein requirements are expressed in terms of DP, diets can be balanced in terms of DP provided data are available regarding the DP concentration of feedstuffs. In France, protein requirements as well as protein content of feeds are expressed in terms of MADC (Matires Azotes Digestibles Cheval). This system was developed to account for differences in digestibility of various sources of protein and availability of protein digestion products in the small versus large intestine. Estimates of true ileal digestibility of protein has been made for hays, concentrates and mixed diets (Martin-Rosset et al., 1994). This system provides insight into considering available protein for the horse rather than CP. A number of studies have measured digestible protein (DP) values for a limited number of feeds in horses, but as yet, insufficient data are available for the horse for DP to be the standard expression for describing either the protein requirements of horses or the protein quality of the wide range of feedstuffs given to them. Not all CP is available to the animal and adjusting CP content for protein that is not available to the horse could allow estimation of digestible or available protein (AP). Available protein can be estimated by subtracting non-protein nitrogen (NPN) from acid detergent insoluble nitrogen (ADIN; which represents bound protein) from CP values for the feeds. Thus, AP is a calculated estimate of protein that may be available to the horse while DP is based on whole tract digestibility studies that measured the amount of protein apparently digested in vivo. Studies that evaluated protein digestibility in the horse were used to evaluate the concept of AP (Hintz and Schryver, 1972; Reitnour and Salsbury, 1972; Reitnour and Salsbury, 1975; Glade et al., 1985; Gibbs et al., 1988; Farley et al., 1995; Gibbs et al., 1996; Crozier et al., 1997; LaCasha et al., 1999). Digestible protein intake was calculated using the digestibility for protein observed in the respective study. The AP of the diet in each study and the respective AP intake was then calculated using the diet and intake data provided in the study. The intake of AP was compared to the DP intake. There was a high correlation between the two variables (r2 = 0.91). Additional research is required to evaluate this concept in vivo but it could be the next step in progressing from CP to DP in terms of expressing protein requirements for horses.

167

Comparisons of AP in various feedstuffs can assist in selecting better sources of protein for the horse (Table 3). Table 3. Select composition of various feedstuffs for calculation of available protein (AP) 1. Feedstuff Bermudagrass Oats CP, % 13.7 13.2 ADFIP, % 1.2 0.3 NPN, % 5.0 8.6 AP, % 7.5 4.3 g CP in 1 kg 137 132 499 518 g AP in 1 kg 75 43 383 187

Soybean Meal 49.9 0.4 11.2 38.3 Peanut Meal 51.8 1.1 32.0 18.7 ADFIP, percentage of a feedstuff that is acid detergent insoluble protein 1 NRC, 2001 Amino Acid Profiles

Nitrogen balance studies have resulted in various estimates of the amounts of nitrogen needed to maintain zero nitrogen balance. The differences may be explained by differences in digestibility. Therefore, 100 g of CP from SBM may be different from 100 g of CP from fishmeal nutritionally to the horse, because of differences in the amino acid profile and digestibility of the protein source. Data from various studies found zero nitrogen balance for fishmeal, SBM and corn gluten meal at intakes of 0.57, 0.8 and 1.18 g CP/kg BW/day, respectively. Such results suggest that different protein sources have varying foregut digestibilities and/or amino acid profiles, and thus differ in their availability to the horse. Based on the previous figures, 285 g, 400 g and 590 g CP would need to be provided to the 500 kg horse for zero nitrogen balance utilizing fishmeal, soybean meal and corn gluten meal respectively. These differences also result in different lysine intakes due to the various lysine concentrations in the feedstuffs. Fishmeal, with a lysine concentration of 5.2%, would provide 14.8 g lysine with 285 g total intake while corn gluten meal at an intake of 590 g would only provide 6.5 g lysine due to its low lysine concentration (1.1%). The use of higher quality protein sources can provide higher quantities of amino acids with lower overall intake. This is illustrated in a study evaluating threonine as a potentially limiting amino acid for yearling horses (Graham et al., 1994). Three diets were fed with similar CP concentrations (12% CP) with varied lysine and threonine concentrations. Improved growth and reduced serum urea nitrogen concentrations were reported in yearlings fed 45 g of lysine (127 mg/kg BW/day) and 39 g of threonine (110 mg/kg BW/day) per day compared to yearlings receiving no amino acid supplementation or supplemental lysine of 42 g/day (116 mg/kg BW/day). This study supports the idea that improvement of protein quality can allow reduction of the overall concentration of CP in the diet. Similar results have been observed in studies that improved the quality of the protein sources with lysine supplementation (Borton et al., 1973; Ott et al., 1981). Improving amino acid profiles for improved protein quality is similar to the concept of ideal protein. Ideal protein is a concept that states that amino acids need to be provided in adequate quantities as well as in the correct proportions to one another. This concept has been evaluated in swine through nitrogen balance studies. The resulting ideal protein for swine is also closely correlated to the major end product of amino acids in the body muscle tissue. Ideal protein for horses has not been studied however, one study has evaluated the amino acid profile of muscle tissue in horses and developed ideal protein ratios of amino acids based on this data (Bryden, 1991). In a study utilizing young and aged exercising horses, dietary protein amounts were fed to provide adequate amounts of dietary CP (based on 1989 NRC) or supplementary lysine and threonine at an amount recommended for growth in an effort to evaluate the horses ability to build and maintain muscle mass. The study reported lower plasma urea nitrogen and 3MH concentrations as well as greater plasma creatinine and subjective muscle mass scores for horses fed the amino acid fortified diets. The lysine intake of the control group was above the current 168

recommendation for lysine for exercising horses (Graham-Thiers et al., 2005). The results suggest that the balance of amino acids was improved with supplementation. The supplemented diet more closely resembled the ideal protein ratios suggested by Bryden (1991) compared to the control diet. Recognizing differences in protein availability as well as amino acid profiles can help us make better decisions regarding protein sources for our horses in various functions. Additional research is needed for horses in the areas of protein digestibility as well as protein availability from various feed sources. This could allow us to move to digestible protein as the way requirements are expressed rather than CP and would allow the feeds to be placed on a basis to allow for better comparison. Specific amino acid requirements as well as the correct balance of these amino acids (ideal protein) deserve attention in the future as well. References
Borton, A., D.R. Anderson and S. Lyford. 1973. Studies of protein quality and quantity in the early weaned foal. Page 19. In: Proc. 3rd Equine Nutr. Phys. Soc. Symp., Gainesville, FL. Boyer, J., N. Cymbaluk, B. Kyle, D. Brown and H. Hintz. 1999. Nitrogen metabolism in pregnant mares fed grass hays containing different concentrations of protein. J. Anim. Sci. 77(Suppl. 1):202. Breuer, L.H., and D.L. Golden. 1971. Lysine requirement of the immature equine. J. Anim. Sci. 33:227. Bryden, W.L. 1991. Amino acid requirements of horses estimated from tissue composition. Page 53. In: Proc. Nutr. Soc. Aust. Crozier, J.A., V.G. Allen, N.E. Jack, J.P. Fontenot and M.A. Cochran. 1997. Digestibility, apparent mineral absorption, and voluntary intake by horses fed alfalfa, tall fescue, and Caucasian bluestem. J. Anim. Sci. 75(6):1651-8. de Almeida, F.Q., S. de Campos Valdares Filo, J.L. Donzele, J.F.C. da Silva, M.I. Leao, P.R. Cecon and C. de Queiroz. 1998. Apparent and true prececal and total digestibility of protein in diets with different protein levels in equines. R. Bras. Zootec. 27(3):521-529. Doreau, M., S. Boulot, W. Martin-Rosset and J. Robelin. 1986. Relationship between nutrient intake, growth and body composition of the nursing foal. Reproduction Nutr. Dev. 26:683-690. Doreau, M., S. Boulot, J.P. Barlet and P.Patureau-Mirand. 1990. Yield and composition of milk from lactating mares: Effect of lactation stage and individual differences. J. Dairy Res. 57:449-454. Farley, E.B., G.D. Potter, P.G. Gibbs, J. Schumacher and M. Murray-Gerzik. 1995. Digestion of soybean meal protein in the equine small and large intestine at varying levels of intake. J. Equine Vet. Sci. 15(8):391-397. Freeman, D.W., G.D. Potter, G.T. Schelling and J.L. Kreider. 1988. Nitrogen metabolism in mature horses at varying levels of work. J. Anim. Sci. 66:407-412. Gibbs, P., G.D. Potter, G.T.Schelling, J.L.Kreider and C.J. Boyd. 1996. The significance of small vs large intestinal digestion of cereal grain and oilseed protein in the equine. J.EquineVet.Sci.16(2):60-65. Gibbs, P.G., G.D. Potter, G.T. Schelling and C.J. Boyd. 1988. Digestion of hay protein in different segments of the equine digestive tract. J. Anim. Sci. 66:400-406. Glade, M.J., D. Beller, J. Bergen, D. Berry, E. Blonder, J. Bradley, M. Cupelo and J. Dallas. 1985. Dietary protein in excess of requirements inhibits renal calcium and phosphorus reabsorption in young horses. Nutr. Rep. Int. 31: 649-659. Graham, P.M., E.A. Ott, J.H. Brendemuhl and S.H. TenBroeck. 1994. The effect of supplemental lysine and threonine on growth and development of yearling horses. J. Anim. Sci. 72:380-386. Graham-Thiers, P.M., and D.S. Kronfeld. 2005. Amino acid supplementation improves muscle mass in aged horses. J. Anim. Sci. 83(12):2783-8. Harper, O.F., and G.W. Vander Noot. 1974. Protein requirements of mature maintenance horses. J. Anim. Sci. 62:183(abstract). Hintz, H.F., and H.F. Schryver. 1972. Nitrogen utilization in ponies. J. Anim. Sci. 34:592-595. LaCasha, P.A., H.A. Brady, V.G. Allen, C.R. Richardson and K.R. Pond. 1999. Voluntary intake, digestibility, and subsequent selection of Matua bromegrass, coastal bermudagrass, and alfalfa hays by yearling horses. J. Anim. Sci. 77(10):2766-73.

169

Mariani, P., A. Summer, F. Martuzzi, P. Formaggioni, A. Sabbioni and A.L. Catalano. 2001. Physiochemical properties, gross composition, energy value and nitrogen fractions of Haflinger nursing mare milk throughout 6 lactation months. Anim. Res. 50:415-425. Martin, R.G., N.P. McMeniman and K.F. Dowsett. 1992. Milk and water intakes of foals sucking grazing mares. Equine Vet. J. 24(4):295-299. Martin-Rosset, W., M.Vermorel, M. Doreau, J.L. Tisserand and J. Andrieu. 1994. French horse feed evaluation systems and recommended allowances for energy and protein. Livestock Prod. Sci. 40:37-56. Meyer, H. 1985. Investigations to determine endogenous faecal and renal N losses in horses. Page 68. In: Proc. 9th Eq. Nutr. Phys. Soc. Symp. East Lansing, MI. NRC (National Research Council). 1989. Nutrient Requirements of Horses, 5th ed. Washington, DC. National Academy Press. NRC. 2001. Nutrient Requirements of Dairy Cattle, 7th ed. Washington, DC. National Academy Press. NRC. 2007. Nutrient Requirements of Horses, 6th ed. Washington, DC. National Academy Press. Olsman, A.F.S., W.L. Jansen, M.M.Sloet van Oldruttenborgh-Oosterbaan and A.C. Beynen. 2003. Assessment of the minimum protein requirement of adult ponies. J. Anim. Physiol. Anim. Nutr. 87:205-212. Ott, E.A., R.L. Asquith, J.P. Feaster and F.G. Martin. 1979. Influence of protein level and quality on growth and development of yearling foals. J. Anim. Sci. 49:620-626. Ott, E.A., R.L. Asquith and J.P. Feaster. 1981. Lysine supplementation of diets for yearling horses. J. Anim. Sci. 53:1496-1503. Ott, E.A., and R.L. Asquith. 1986. Influence of level of feeding and nutrient content of the concentrate on growth and development of yearling horses. J. Anim. Sci. 62:290-299. Ott, E.A., and J. Kivipelto. 2002. Growth and development of yearling horses fed either alfalfa or coastal bermudagrass: hay and a concentrate formulated for bermudagrass hay. J. Eq. Vet. Sci. 22(7):311-319. Patterson, P.H., C.N. Coon and I.M. Hughes. 1985. Protein requirements of mature working horses. J. Anim. Sci. 61:187-196. Platt, H. 1984. Growth of the equine foetus. Equine Vet. J. 16(4): 247-252. Potter, G.D., P.G. Gibbs, R.G. Haley and C. Klendshoj. 1992. Digestion of protein in the small and large intestines of equines fed mixed diets. Pferdeheikunde 140-143. Reitnour, C.M., and R.L. Salsbury. 1972. Digestion and utilization of cecally infused protein by the equine. J. Anim. Sci. 35(6):1190-3. Reitnour, C.M., and R.L. Salsbury. 1975. Effect of oral or cecal administration of protein supplements on equine plasma amino acids. Br. Vet. J. 131:466-472. Reitnour, C.M., and R.L. Salsbury. 1976. Utilization of proteins by the equine species. Am. J. Vet. Res. 37:1065-1067. Slade, L.M., D.W. Robinson and K.E. Casey. 1970. Nitrogen metabolism in nonruminant herbivores. I. The influence of nonprotein nitrogen and protein quality on the nitrogen retention of adult mares. J. Anim. Sci. 30(5):753-760. Smolders, E.A., N.G. Van Der Veen and A. Polanen. 1990. Composition of horse milk during the suckling period. Livestock Prod. Sci. 25:163-171. van Niekerk, F.E., and C.H. van Niekerk. 1997. The effect of dietary protein on reproduction in the mare: I. The composition and evaluation of digestibility of dietary protein from different sources. J. S. Afr. Vet. Assoc. 68(3):78-80. Wickens, C.L., P.K. Ku and N.L. Trottier. 2002. An ideal protein for the lactating mare. J. Anim. Sci. (Suppl. 1):155. Wickens, C.L., J. Moore, J. Shelle, C. Skelly, H.M. Clayton and N.L. Trottier. 2003. Effect of exercise on dietary protein requirement of the Arabian horse. Page 128. In: Proc. 18th Eq. Nutr. Phys. Soc. Symp., East Lansing, MI. Wickens, C.L., J. Moore, C. Wolf, C. Skelley and N. Trottier. 2005. 3-methylhistidine as a response criterium to estimate dietary protein requirement of the exercising horse. Page 205. In: Proc. 19th Equine Sci. Soc., Tucson, AZ.

170

USING THE NRC TO MANAGE HORSE NUTRITION (AN OVERVIEW OF MANAGEMENT ISSUES IN THE NUTRIENT REQUIREMENTS OF HORSES, SIXTH REVISED EDITION) David W. Freeman Oklahoma State University Department of Animal Science Stillwater, OK 74078 Phone: 405-744-6058 FAX: 405-744-7390 Email: david.freeman@okstate.edu Summary This paper provides an overview of some of the management topics that are discussed in relative detail in the newly revised Nutrient Requirements of Horses (NRC, 2007). The NRC, 2007 has been revised for the sixth time in its publication history. The primary goal of the NRC revision was to provide updated recommendations on nutrient requirements of horses of different weights and productive states. Chapters addressing nutrients requirements provide a review of literature and logic leading to derivation of formulas developed to estimate requirement levels. Information for feeding management can be extracted from these chapters; however, the main chapters dedicated to management issues are those that discuss specific feedstuffs and feed additives, feeding behavior, and those covering unique aspects of nutrition or the feeding of donkeys and wild equid in captivity. One significant area of interest with feeding management of horses relates to managing carbohydrate intake, especially the feeding of starch and nonstructural carbohydrates. Recommendations are to limit single meal feeding of starch to a maximum of 2.0 g to 4.0 g starch/kg BW. However, influences of variation from horse differences, feedstuffs and intake schedules alter the utilization of carbohydrates; hence, there is need to consider these and other potential sources of variation when designing feeding management plans. Similarly, intake behavior factors into successful feeding management. As such, a significant review of feeding behavior is included in the NRC, 2007. Additionally, feed analysis and ration evaluation are important tools relating to feeding management, and both subjects have expanded reviews in this edition. Several general feeding management guidelines are provided, including body weight management. Specific chapters on unique aspects of feeding management, and the feeding of donkeys and wild equids in captivity also are new additions to the NRC, 2007. Introduction The Sixth Revised Edition of the National Research Council publication entitled Nutrient Requirements of Horses (NRC, 2007) is in the printing process (as of January, 2007). The publication can be pre-ordered from the National Academies Press website: www.nap.edu. It has been 18 years since the last revision. Chapters are organized in nutrient categories (energy, protein, mineral, vitamin and water), with supportive chapters reviewing literature specific to nutrient type (fats, carbohydrate), feedstuffs (feeds and feed processing, feed additives) and feed composition. Further, the revision provides in-depth reviews of topics more typically associated with feeding management (feeding behavior, feed analysis, ration evaluation) or topics on particular areas of interest (unique aspects, donkeys and other equids). Tabular values identifying nutrient requirements of horses with different weights, ages or states of production, and tables identifying nutrient content of selected feedstuffs, mineral supplements, and mares milk are included in the NRC, 2007. Also included is a chapter with summary formulas used to estimate nutrient requirements.

Proceedings of the 5th Mid-Atlantic Nutrition Conference. 2007. Zimmermann, N.G., ed., University of Maryland, College Park, MD 20742

171

The primary goal of the NRC revision was to provide updated recommendations on nutrient requirements of horses of different weights and productive states. Chapters addressing nutrients requirements provide a review of literature and logic leading to derivation of formulas developed to estimate requirement levels. Information for feeding management can be extracted from these chapters; however, the main chapters dedicated to management issues are those that discuss specific feedstuffs and feed additives, feeding behavior, and those covering unique aspects of nutrition or the feeding of donkeys and wild equid in captivity. The goal of feeding management is to provide nutrients that efficiently maintain a horses body and well-being, and support functions related to growth, production, and work. This paper provides an overview of some of the nutrient management topics discussed in relative detail in the sixth edition of The Nutrient Requirements for Horses (NRC, 2007). Management of Feeding Carbohydrates Perhaps the most rapidly developing research area of horse nutrition involves study of feeding various carbohydrate fractions. Currently, the only carbohydrate fraction listed on feed tags is crude fiber, one of the components analyzed when performing proximate analysis. Animal nutritionists have for many years expanded their analysis of fiber to include neutral detergent and acid detergent fiber, partitioning the less digestible fiber from those fibrous compounds that typically are more digestible. The nonfibrous carbohydrates, which are more readily digested throughout the equine digestive tract as compared with fibrous carbohydrates, lack a satisfactory system of classification in the equine as neither the ruminant (Van Soest et al., 1991) nor human (Englyst and Cummings, 1990) systems of classification fit well with the digestive physiology or carbohydrate metabolism of horses (Hoffman, 2004). As such, current day researchers are suggesting different partitioning schemes and publishing analytical values previously unresearched in horses, i.e. hydrolysable carbohydrates, fermentable carbohydrates, and slowly fermented carbohydrates (Hoffman, 2004). Reports also are publishing values for specific classes of compounds with inference to horse health, i.e. starch, fructans and simple sugars (Hoffman et al., 2003; Pollit et al., 2003; Potter et al., 1992a). This influx of terminology, partitioning schemes and research makes ration evaluation more complex, and is one of the most needed areas of continued research in horse nutrition. A significant number of recently published studies have centered on the digestion of starch. Many horses are meal-fed rations with significant levels of starch. Total tract starch digestibility ranges typically between 87 percent to nearly 100 percent (Potter et al., 1992a). Starch that is not digested in the small intestine is readily digested by microbes in the hindgut. A large amount of starch presented at one time to the hindgut is thought to increase digestive upset and result in lactate acidosis, colic and laminitis (Garner et al., 1977; Medina et al., 2002; Pollitt et al., 2003; Tinker et al., 1997). Recommendations are to limit single meal feeding of starch to a maximum of 2.0 g to 4.0 g starch/kg BW in order to avoid such disorders (Kienzle et al., 1992; Potter et al., 1992a). The rate and site of digestion of starch is influenced by many factors, including ration intake level, morphology and processing of starch, rate of passage and differences between horses (Cuddeford, 1999; Hussein and Vogedes, 2003; Jose-Cunilleras et al., 2004; Kienzle et al., 1997; Meyer et al., 1993). Several trials have documented the differences in foregut digestibility of starch relative to different feedstuffs and processing methods (Kienzle et al., 1992; Meyer et al, 1993; Potter et al., 1992a). There appears to be significant differences in pre-cecal starch digestion between commonly fed grains, as well as major effects due to the extent and type of processing. The following table is an example of NRC, 2007 reporting of these effects.

172

Table 1. Comparison of Small Intestinal Starch Digestibility of Grains Fed at Moderate Intakes.a,b Crimped Oats Micronized Oats Crimped Sorghum Micronized Sorghum 2.8 59.0 94.5 Mean 2.7 51.3 94.1

Starch Intake (g/kg body 2.8 2.6 2.9 weight/meal) Pre-cecal Starch 48.0 62.3c 36.0d Digestibility (%) Total Tract Starch 94.4 93.8 94.0 Digestibility a As reported in Potter et al. (1992a). b Digestibility reflective of total ration of grain fed with Bermudagrass hay. c,d Values with different superscripts reported different at P< 0.1.

Any effect of improving total tract digestibility of nonstructural carbohydrates or more importantly, altering the site of digestion has potential benefit. However, quantifying the expected benefits of processing for the large variety of feeding plans and routines of horse management is difficult. Grain type, horse differences, processing method, and level of starch and dry matter intake are important influencing factors that are not fully understood. Logically, processing methods that have been shown to improve the site of digestion of starch have significant potential for reducing digestive upset, especially when horses are fed meals containing large amounts of starch that if unprocessed is resistant to small intestinal digestion. Most nutritionists and veterinarians emphasize the feeding of fibrous carbohydrates to meet the energy needs of horses, usually referring to the need for forage-based diets. Even though there are no reports of trials that prove a minimal forage or fiber need, circumstantial evidence suggests that insufficient fiber leads to increased reliance on intake of starch and nonstructural carbohydrates which increases the incidence of hindgut acidosis, colic, gastric ulcers, and behavioral abnormalities (Medina et al., 2002; McGreevy et al., 1995; Murray and Schusser, 1989; Tinker et al., 1997). Review of research literature generally reinforces the traditional management recommendations of emphasizing the harvesting and grazing of forages with relatively lower levels of lignin and cellulose, and higher levels of non-structural carbohydrates, hemicellulose and pectin. The relative distribution of the various fiber fractions of forages differ with environmental conditions, stage of maturity and forage type (Beever et al., 2000; Buxton et al., 1995; Longland et al., 1995). Emphasizing forage-based diets for horses is one method of decreasing the concern of meal feeding diets with significant starch levels. Another method is the incorporation of high fiber feedstuffs into complete feeds. Several byproducts have been the focus of equine nutritional research and are routinely incorporated into equine rations to supply a readily available source of fiber, i.e. soybean hulls (Coverdale et al., 2004) and beet pulp (Harris and Rodiek, 1993). The prevalence of use of byproduct feedstuffs as large components of horse diets, such as rice bran, distillers grains, wheat midds, and soybean hulls, increases the need for more research to be conducted on the byproduct use with different feeding plans, and production and use classes of horses. Feeding Behavior One obvious management concern with feeding behavior is the monitoring and regulation of intake. As nutritionists, veterinarians and horse owners will readily attest, most horses will eat beyond their need to meet energy requirements when given the opportunity. The application of information retrieved from research on feeding behavior is difficult as influences of diet, housing, environment and horse variation complicate the accuracy of estimating expected level of intake. Review of research suggests that mature and young horses appear to have a maximal daily dry matter feed intake ranging

173

from 1.8 percent to 3.2 percent of body weight (Aiken et al., 1989; Crozier et al., 1997; Dulphy et al., 1997a; Dulphy et al., 1997b; Heusner, 1993; Marlow et al. 1983). On a practical basis, estimating maximal voluntary intake is more of a concern when feeding or grazing forages rather than meal-fed grains. Average daily voluntary dry matter intake of forages generally range between 2 percent and 3 percent of body weight (Aiken et al., 1989; Cymbaluk et al., 1989; Guay et al., 2002), although there is an expectation of large variation. There are numerous influences that alter voluntary intake and grazing behavior of forages including plant maturity (Fleurance et al., 2001), availability of plant variety (Archer, 1973), gender and age of horse (Crowell-Davis et al., 1985; Mesochina et al., 2000), season grazed (Salter and Hudson, 1979; Duncan, 1985), availability of pasture (Arnold, 1984) and environmental conditions such as rain (Rogalski, 1975) and snow (Kawai et al., 2004). Understanding feeding behavior, what is expected and what influences to account for, will aid the horse owner in the design of feeding plans and management of feed sources. Perhaps a logical path is to accurately identify nutrient intake, estimate digestibility based on research of the feed or similar feedstuffs, and then weigh the relative impact of factors that influence the uniqueness of supplying nutrients to particular horses. Feed Analysis and Ration Evaluation The needs to understand feed analyses is becoming more important to horse owners as greater varieties of feedstuffs become major components of horse rations, and as different compounds are researched. Additionally, nutritionists and veterinarian consultants will need a heightened level of knowledge of which compounds are included in various analyses. This need is especially important in the area of carbohydrate analysis, as neither the existing ruminant nor human systems fit well with the digestive physiology or intermediary metabolism of the horse. This need is likely to receive heightened number of requests from horse owners. Also, analysis of protein quality will become more routine as research emphasis is being placed on the site of nitrogen digestion (Potter et al., 1992b) digestibility of individual amino acids (deAlmeida et al., 1998), and amino acid requirements (Graham et al., 1994; Ott and Kivipelto, 2002; Wickens et al., 2002). Ideally, all feeds should be analyzed, especially those with newly incorporated feedstuffs with known histories of large variability. Arosemena et al. (1995) related the extent of nutrient variability of several byproduct feedstuffs to values listed in the Nutrient Requirements for Dairy Cattle (NRC, 2001). Comparison of average nutrients of the analyzed byproduct feeds differed more than 20 percent from values in the NRC (2001) table for most nutrients. Largest variations within sources of the sampled feedstuffs tended to be within mineral levels. Others have noted similar variability in the composition between different sources of byproduct feeds (DePeters et al., 2000) and differences in nutrient composition reported by commercial laboratories compared to those reported in NRC feedstuff composition tables (Berger, 1996). Feed analysis, with an understanding of compounds contained in the various components of analysis, is the first step of managing the feeding of nutrients. To that means, the NRC has expanded discussion of various components of analyses, including those compounds not frequently analyzed because of costs or because components have not been studied sufficiently for inclusion in routine analyses. A web-based computer program will be available on the National Academies website that will allow horse owners to obtain recommended nutrient needs. Users are required to input animal parameters such as weight, production state, and level of activity. The need to evaluate nutrient content of rations and relate intake to these recommendations is of major concern for many horse owners. For that reason, several simple examples outlining mathematical procedures for evaluation are provided in the NRC, 2007.

174

General Feeding Management Guidelines Several general feeding management guidelines are provided in the NRC, 2007 that are intended to be a summary of the many management principles mentioned throughout the publication. Emphasis is placed on the need to account for supplies of nutrients from prepared mixes and forages before making adjustments with on-site formulations or supplementation. Feeding by weight rather than volume is attributed as an important aspect of accurate accounting of nutrient intake. Management of meal-feeding is primarily directed to management of starch intake. Management of forage intake emphasizes advantages of use of fiber as a major supplier of energy. Managing body condition through the use of body condition scoring is discussed in connection with weight management. Other Subjects of Potential Interest to Horse Owners The NRC, 2007 contains 340 pages of text, tables and figures. A significant amount of information is provided on composition of feeds and feed additives. A chapter outlining special considerations for donkeys and other equids reviews feeding behavior, nutrient requirements and feeding management of donkeys and wild equids in captivity. In addition to an extensive discussion on the nutrition management of numerous disease conditions, the feeding of nursing and orphan foals and aged horses, and the feeding management of horses in cold or hot weather are covered as part of the chapter on unique aspects. References Aiken, G.E., G.D. Potter, B.E. Conrad and J.W. Evans. 1989. Voluntary intake and digestion of coastal Bermuda grass hay by yearling and mature horses. J. Equine Vet. Sci. 9:262-264. Arosemena, A., E. J. DePeters and J. G. Fadel. 1995. Extent of variability in nutrient composition within selected by-product feedstuffs. Anim. Feed Sci. Tech. 54:103-120. Archer, M. 1973. The species preference of grazing horses. J. Brit. Grassland Soc. 28:123-128. Arnold, G.W. 1984. Comparison of the time budgets and circadian patterns of maintenance activities in sheep, cattle and horses grouped together. Appl. Animal Behav. Sci. 13:19-30. Beever, D.E., N. Offer and M. Gill. 2000. The feeding value of grass and grass products. Chapter 7. In: Grass: Its Production and Utilisation, 3rd ed. Hopkins, A., ed. Oxford, UK. British Grassland Society, Blackwell Science Publications. Berger, L.L. 1996. Variation in the trace mineral content of feedstuffs. Prof. Anim. Sci. 12:1-5. Buxton, D.R., D.R. Mertens and K.J. Moore. 1995. Forage quality for ruminants: plant and animal considerations. Prof. Anim. Sci. 11:121. Coverdale, J.A., J.A. Moore, H.D. Tyler and P.A. Miller-Auwerda. 2004. Soybean hulls as an alternative feed for horses. J. Anim. Sci. 82:1663-1668. Crowell-Davis, S.L., K.A. Houpt and J. Carnevale. 1985. Feeding and drinking behavior of mares and foals with free access to pasture and water. J Anim. Sci. 60:883-889. Crozier, J.A., V.G. Allen, N.E. Jack, J.P. Fontenot and M.A. Cochran. 1997. Digestibility, apparent mineral absorption and voluntary intake by horses fed alfalfa, tall fescue and Caucasian bluestem. J. Anim. Sci. 75:1651-1658. Cuddeford, D. 1999. Starch digestion in the horse. Pages 129-139. In: Advances in Equine Nutrition, Proc. 1998 Equine Nutr. Conf. for Feed Manufacturers. Kentucky Equine Research, Inc. Cymbaluk, N.F., G.I. Christison and D.H. Leach. 1989. Energy uptake and utilization by limit and ad libitum-fed growing horses. J. Anim. Sci. 67:403-413. de Almeida, F.Q., S. de Campos Valdares Filo, J.L. Donzele, J.F.C. da Silva, M.I. Leao, P.R. Cecon and C. de Queiroz. 1998b. Endogenous amino acid composition and true prececal apparent and true digestibility of amino acids in diets for equines. R. Bras. Zootec. 27(3):546-555.

175

DePeters, E.J., J.G. Fadel, M.J. Arana, N. Ohanesian, M.A. Etchebarne, C.A. Hamilton, R.G. Hinders, M.D. Maloney, C.A. Old, T.J. Riordan, H. Perez-Monti and J.W. Pareas. 2000. Variability in the chemical composition of seventeen selected by-product feedstuffs used by the California dairy industry. Prof. Anim. Sci. 16:69-99. Dulphy, J.P., W. Martin-Rosset, H. Dubroeucq, J.M. Ballet, A. Detour and M. Jailler. 1997a. Compared feeding patterns in ad libitum intake of dry forages by horses and sheep. Livestock Prod. Sci. 52:4956. Dulphy, J.P., W. Martin-Rosset, H. Dubroeucq and M. Jailler. 1997b. Evaluation of voluntary intake of forage trough-fed to light horses. Comparison with sheep. Livestock Prod. Sci. 52:97-104. Duncan, P. 1985. Time-budgets of Camargue horses. III. Environmental influences. Behavior 92:188-208. Englyst, H., and J. Cummings. 1990. Dietary fibre and starch: Definition, classification and measurement. Pages 3-26. In: Dietary Fibre PerspectivesReviews and Bibliography. Leeds, A.R. (ed.). John Libbey and Co., London. Fleurance, G., P. Duncan and B. Mallevaud. 2001. Daily intake and the selection of feeding sites by horses in heterogeneous wet grasslands. Anim. Res. 50:149-156. Garner, H.E., D.P. Hutcheson, J.R. Coffman and A.W. Hahn. 1977. Lactic acidosis. A factor associated with equine laminitis. J. Anim. Sci. 45(5):1037-1041. Graham, P.M., E.A. Ott, J.H. Brendemuhl and S.H. TenBroeck. 1994. The effect of supplemental lysine and threonine on growth and development of yearling horses. J. Anim. Sci. 72:380-386. Guay, K.A., H.A. Brady, V.G. Allen, K.R. Pond, D.B. Wester, L.A. Janecka and N.L. Heninger. 2002. Matua bromegrass hay for mares in gestation and lactation. J. Anim. Sci. 80:2960-2966. Harris, D.M., and A.V. Rodiek. 1993. Dry matter digestibility of diets containing beet pulp fed to horses. Pages 100-101. In: Proc. 13th Equine Nutr. Physiol. Symp., University of Florida, Gainesville, FL. Heusner, G.L. 1993. Ad libitum feeding of mature horses to achieve rapid weight gain. Pages 86-87. In: Proc. 13th Equine Nutr. Physiol. Symp., Gainesville, FL. Hoffman, R.M. 2004. Carbohydrate in horse nutrition. Pages 21-37. In: Proc. Conference on Equine Nutrition Research, Texas A&M University, Equine Science Section, Department of Animal Science. Hoffman, R.M., J.A. Wilson, D.S. Kronfeld, W.L. Cooper, L.A. Lawrence, D. Sklan and P.A. Harris. 2001. Hydrolyzable carbohydrates in pasture, hay and horse feeds: direct assay and seasonal variation. J. Anim. Sci. 79:500-506. Hussein, H.S., and L.A. Vogedes. 2003. Review: forage nutritional value for equine as affected by forage species and cereal grain supplementation. Prof. Anim. Sci. 19:388-397. Jose-Cunilleras, E., L.E. Taylor and K.W. Hinchcliff. 2004. Glycemic index of cracked corn, oat groats and rolled barley in horses. J. Anim. Sci. 82:2623-2629. Kawai, M., H. Hisano, Y. Yabu, N. Yabu and S. Matsuoka. 2004a. Effects of fallen snow on the voluntary intake and grazing behavior of Hokkaido native horses in winter woodland with underlying Sasa Senanensis. Anim. Sci. J. 75:435-440. Kienzle, E., S. Radicke, S. Wilke, E. Landes and H. Meyer. 1992. Preileal starch digestion in relation to source and preparation of starch. Pages 103-107. In: Eur. Conf. on Nutrition for the Horse, Hannover, Germany. Kienzle, E., J. Pohlenz and S. Radicke. 1997. Morphology of starch digestion in the horse. JAVMA 44:207-221. Longland, A.C., R. Pilgrim and I.J. Jones. 1995. Comparison of oven drying vs. freeze drying on the analysis of non-starch polysaccharides in gramminaceous and leguminous forages. Proc. Brit. Soc. Anim. Sci.:60. Marlow, C.H.B., E.M. van Tonder, F.C. Hayward, S.S. van der Merwe and L.E.G. Price. 1983. A report on the consumption, composition and nutritional adequacy of a mixture of lush green perennial ryegrass (Lolium perenne) and cocksfoot (Dactylis glomerata) fed ad libitum to thoroughbred mares. J. South African Vet. Assoc. 54(3): 155-157.

176

McGreevy, P.D., P.J. Cripps, N.P. French, L.E. Green and C.J. Nicol. 1995. Management factors associated with stereotypic and redirected behavior in the Thoroughbred horse. Equine Vet. J. 27 (2):86-91. Medina, B., I.D. Girard, E. Jacotot and V. Julliand. 2002. Effect of a preparation of Saccharomyces cerevisiae on microbial profiles and fermentation patterns in the large intestine of horses fed a high fiber or a high starch diet. J. Anim. Sci. 80:2600-2609. Mesochina, P.M., J.L Peyraud, P. Duncan, D. Micol and C. Tillaud-Geyl. 2000. Grass intake by growing horses at pasture: a test on the effect of the horses age and sward biomass. Ann. Zootech. 49:505-515. Meyer, H., S. Radicke, E. Kienzle, S. Wilke and D. Kleffken. 1993. Investigations on preileal digestion of oats, corn and barley starch in relation to grain processing. Pages 92-97. In: Proc. 13th Eq. Nutr. Phys. Symp., University of Florida, Gainesville. Murray, M.J., and G. Shusser. 1989. Application of gastric pH-metry in horses: Measurement of 24 h gastric pH in horses fed, fasted and treated with ranitidine. J. Vet. Internal Med. 6:133. NRC (National Research Council). 2007. Nutrient Requirements of Horses, 6th rev. ed. Washington, DC. National Academy Press. NRC (National Research Council). 2001. Nutrient Requirements of Dairy Cattle, 7th rev. ed. Washington, DC. National Academy Press. Ott, E.A., and J. Kivipelto. 2002. Growth and development of yearling horses fed either alfalfa or coastal Bermuda grass: hay and a concentrate formulated for Bermuda grass hay. J. Equine Vet. Sci. 22(7):311-319. Pollitt, C.C., M. Kyaw-Tanner, K.R. French, A.W. Van Eps, J.K. Hendrikz and M. Daradka. 2003. Equine Laminitis. Pages 103-115. In: 49th Annual Convention of the American Association of Equine Practitioners, New Orleans, LA. American Association of Equine Practitioners. Potter, G.D., F.F. Arnold, D.D. Householder, D.H. Hansen and K.M. Brown. 1992a. Digestion of starch in the small or large intestine of the equine. Pages 107-111. In: European Conference on Nutrition for the Horse, Hanover, Germany. Potter, G.D., P.G. Gibbs, R.G. Haley and C. Klendshoj. 1992b. Digestion of protein in the small and large intestines of equines fed mixed diets. Pferdeheikunde 1:140-143. Rogalski, M. 1977. Behaviour of animals on pasture. Roczniki Akademii Rolniczej w Poznaniu, Rozprawny Naukowe. 78:41pp. Salter, R.E., and R.J. Hudson. 1979. Feeding ecology of feral horses in western Alberta. J. Range Management 32:221-225. Tinker, M.K., N.A. White, P. Lessard, C.D. Thatcher, K.D. Pelzer, B. Davis and D.K. Carmel. 1997. A prospective study of equine colic risk factors. Equine Vet. J. 29: 454-458. Van Soest, P.J., J.B. Robertson and B.A. Lewis. 1991. Methods for dietary fiber, neutral detergent fiber, and nonstarch polysaccharides in relation to animal nutrition. J. Dairy Sci. 74:3583-3597. Wickens, C.L., P.K. Ku and N.L. Trottier. 2002. An ideal protein for the lactating mare. J. Anim. Sci. (Suppl. 1):155.

177

You might also like