Proceedings - Corona Range and Livestock Research
Transcription
Proceedings - Corona Range and Livestock Research
PROCEEDINGS OF THE 4th Grazing Livestock Nutrition Conference July 9-10, 2010 Estes Park Conference Center Estes Park, Colorado USA Organized by: W 1012 (Formerly WERA 110) “Improving Ruminant Use of Forages in Sustainable Production Systems for the Western United States” Untitled-1 1 6/5/2003, 2:42 PM 4th GRAZING LIVESTOCK NUTRITION CONFERENCE July 9-10, 2010 Estes Park Conference Center, Estes Park, CO Organized by: W 1012 (Formerly WERA – 110) “Improving Ruminant Use Of Forages in Sustainable Production Systems for the Western United States” SCHEDULE THURSDAY, JULY 8, 2010 6:00 – 8:00 p.m. Registration and Poster Set-Up FRIDAY, JULY 9, 2010 8:00 – 8:30 a.m. Registration Opening Remarks - Richard Waterman 8:30 – 8:40 a.m. Welcome and Introductions Moderator - Mark Petersen 8:40 – 8:45 a.m. History 8:45 – 9:30 a.m. Issues in Grazing Livestock Nutrition Dr. Tim DelCurto - Oregon State University, Union, OR Dr. Ken Olson - South Dakota State University, Rapid City, SD SESSION I: RUMINAL MICROBIAL ECOLOGY AS INFLUENCED BY THE NUTRITIONAL ENVIRONMENT OF GRAZING RUMINANTS Moderator - Doug Tolleson 9:30 – 10:15 a.m. Emerging Methods in Rumen Microbiology Dr. Bryan White - University of Illinois, Urbana, IL 10:15 – 10:45 a.m. BREAK 10:45 – 11:30 a.m. Application of Rumen Microbiological Techniques to the Grazing Ruminant Dr. Shanna Ivey - New Mexico State University, Las Cruces, NM 11:30 a.m. – 1:00 p.m. LUNCH SESSION II: PERSPECTIVES ON INTAKE IN GRAZING STUDIES AND IN PRACTICE KEYNOTE SPEAKER Moderator - Jim Sprinkle 1:00 – 2:00 p.m. Assessment of Intake for Grazing Ruminants Dr. Hugh Dove - CSIRO Plant Industry, Canberra, Australia 2:00 – 2:15 p.m. Geospatial Methods and Data Analysis for Assessing Distribution of Grazing Livestock Dr. Dean Anderson - USDA-ARS, Las Cruces, NM SESSION III: IMPACT OF NUTRITIENT BALANCE ON BIOLOGICAL EFFICIENCY AND LIFETIME PRODUCTIVITY OF GRAZING LIVESTOCK Moderator - Jim Carpenter 2:45 – 3:30 p.m. Metabolic Signals of the Beef Cow in Negative Energy Balance Dr. Richard Waterman - USDA –ARS Miles City, MT Dr. Ron Butler - Cornell University, Ithaca, NY 3:30 – 3:45 p.m. BREAK 3:45 – 4:30 p.m. Maternal Plane of Nutrition Impacts on Newborn/Offspring Dr. Joel Caton - North Dakota State University, Fargo, ND Dr. Bret Hess - University of Wyoming, Laramie, WY 4:30 – 5:15 p.m. Excess Nutrients Leading to Imbalance, High-Quality Forages Dr. Monty Kerley - University of Missouri, Columbia, MO 5:15 – 5:30 p.m. Wrap-up and announcements 5:30 – 7:00 p.m. Social and poster session (Presenters attending posters) SATURDAY, JULY 10, 2010 8:00 – 8:15 a.m. Meet and Greet 8:15 – 8:30 a.m. Welcome and announcements SESSION IV: APPLICATION OF REQUIREMENT SYSTEMS FOR GRAZING LIVESTOCK; ASSESSING NRC AND CSIRO SYSTEMS FOR GRAZING LIVESTOCK Moderator - Sergio A. Soto-Navarro 8:30 – 9:30 a.m. Application of Nutrient Requirement Schemes to Grazing Animals Dr. Stuart McLennan - Queensland Primary Industries and Fisheries Animal Research Institute, Moorooka Q, Australia Dr. Hugh Dove - CSIRO Plant Industry, Canberra, Australia Prof. Dennis Poppi - University of Queensland, Gatton, Australia SESSION V: SUPPLEMENTATION STRATEGIES TO ACHIEVE BIOLOGICAL AND ECONOMIC EFFICIENCY Moderator - Eric Scholljegerdes 9:30 – 10:15 a.m. Strategic Supplementation to Correct for Nutrient Imbalances Dr. Greg Lardy - North Dakota State University, Fargo, ND Dr. Rachel Endecott - Montana State University, Miles City, MT 10:15 – 10:45 a.m. BREAK 10:45 – 11:30 a.m. Impact of Nutritional Decisions on Net Returns in an Era of Increasing Costs for Feedstuffs Dr. Allen Torell - New Mexico State University, Las Cruces, NM Dr. Neil R. Rimbey – University of Idaho, Caldwell, ID SESSION VI : A SUMMARY OF ADVANCES AND EMERGING PRACTICES IN GRAZING LIVESTOCK NUTRITION Moderator - Richard Waterman 11:30 – 12:30 p.m. Challenges to Predicting Productivity of Grazing Ruminants - Where to Now? Dr. Mark Petersen - USDA-ARS Miles City, MT 12:30 – 12:45 p.m. Closing Remarks Recognition: -Sponsors -Editorial committee -Organizing committee 12:45 – 1:00 p.m. Remove Posters Foreword Welcome to the Fourth Grazing Livestock Nutrition Conference. The first call-to-order occurred in July 1987 at Jackson Hole, Wyoming. That meeting was dedicated to Drs. Lorin Harris and C. Wayne Cook who inspired the science of Range Livestock Nutrition in the Western United States. Therefore, we would like to dedicate this fourth conference to members of the original planning committee, Dr. D. C. Clanton, Professor Emeritus, University of Nebraska, Dr. J. E. Wallace, Professor Emeritus, New Mexico State University, and Dr. F. Hinds, Professor Emeritus, University of Wyoming. Much of the discussions at this conference are outcomes of their research. We are indebted to them for their foresight and example. The inspiration for the Grazing Livestock Nutrition Conference is a desire to facilitate a forum where researchers, practitioners and students meet, explore advances, and learn the science of grazing animal nutrition. Estes Park is a perfect location that is appealing, with a beautiful outdoor venue, to encourage participants coming together for formal and informal discussions. Much of the conference agenda is related to semi-arid and arid extensive rangelands, although topics of importance to improved pasture settings are included. Another important consideration in the planning of this meeting includes the realization that livestock grazing occurs around the globe primarily in extensive grazing scenarios; this recognition has prompted an international aspect to this meeting. Highly respected international researchers have been key congress speakers presenting challenging theories and intriguing experimental results. Their contribution in the past and this year is highly valued. The Livestock Grazing Nutrition Conference occurs infrequently with 13 years passing since the adjournment of the third conference. There have been many technological advances during that period. The conference papers, abstracts, and discussion reflect some of these developments. This 4th meeting was planned and organized by members of the Multistate Research Project, W1012. We hope the meeting and proceedings are useful. Sincerely, Planning Committee Sponsors Western Section American Society of Animal Science USDA – NIFA* Alpharma Animal Health Cargill Kahne Animal Health Agricultural Experiment Stations (Arizona, Colorado, Montana, and Wyoming) Conference Organizing Committee Richard C. Waterman, USDA – ARS Fort Keogh LARRL, Chairmen Bret W. Hess, University of Wyoming Jack C. Whittier, Colorado State University Mark K. Petersen, USDA – ARS Fort Keogh LARRL Cover Designers Jack C. Whittier, Colorado State University Richard C. Waterman, USDA – ARS Fort Keogh LARRL Cover photo provided by Colorado State University Photographic Services CITATION FORMAT†: Author(s). 2010. Paper title. Pages __–__ in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J.G.P. Bowman, and R.C. Waterman eds. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. *This project was supported by Agriculture and Food Research Initiative grant no. 2010-65206-20719 from the USDA National Institute of Food and Agriculture. † The editorial committee wishes to express their sincere gratitude to those individuals who donated their time and effort to provide professional reviews of manuscripts and abstracts published in this proceeding. Grazing Livestock Nutrition Conference Table of Contents ISSUES IN GRAZING LIVESTOCK NUTRITION T. DelCurto and K. C. Olson ..................................................................................................................... 1 EMERGING METHODS IN RUMEN MICROBIOLOGY J. M. Brulc, C. J. Yeoman, K. E. Nelson, and B. A. White ...................................................................... 10 PRACTICAL APPLICATION OF MODERN RUMEN MICROBIOLOGICAL TECHNIQUES TO GRAZING RUMINANTS S. L. Lodge-Ivey ....................................................................................................................................... 24 KEYNOTE: ASSESSMENT OF INTAKE AND DIET COMPOSITION OF GRAZING LIVESTOCK H. Dove ..................................................................................................................................................... 31 GEOSPATIAL METHODS AND DATA ANALYSIS FOR ASSESSING DISTRIBUTION OF GRAZING LIVESTOCK D. M. Anderson......................................................................................................................................... 57 METABOLIC SIGNALS OF THE BEEF COW IN NEGATIVE ENERGY BALANCE R. C. Waterman and W. R. Butler ............................................................................................................ 93 MATERNAL PLANE OF NUTRITION: IMPACTS ON FETAL OUTCOMES AND POSTNATAL OFFSPRING RESPONSES J. S. Caton and B. W. Hess ..................................................................................................................... 104 POTENTIAL FOR NUTRITIONAL IMBALANCE IN HIGH-QUALITY FORAGES M. Kerley ................................................................................................................................................ 125 APPLICATION OF NUTRIENT REQUIREMENT SCHEMES TO GRAZING ANIMALS H. Dove, S. R. McLennan, and D. P. Poppi ............................................................................................ 133 STRATEGIC SUPPLEMENTATION TO CORRECT FOR NUTRIENT IMBALANCES G. P. Lardy and R. L. Endecott ............................................................................................................... 152 ECONOMICALLY EFFICIENT SUPPLEMENTAL FEEDING AND THE IMPACT OF NUTRITIONAL DECISIONS ON NET RANCH RETURNS L. Allen Torell and Neil R. Rimbey........................................................................................................ 170 CHALLENGES TO PREDICTING PRODUCTIVITY OF GRAZING RUMINANTS: WHERE TO NOW? M. K. Petersen, J. T. Mulliniks, A. J. Roberts, and R. C. Waterman ...................................................... 180 Abstracts 1 EFFECTS OF RESTRICTING TIME AT PASTURE ON FEEDING STATION BEHAVIOR OF CATTLE DURING THE FIRST GRAZING SESSION OF THE DAY. P. Gregorini, K. McLeod, C. Clark, C. Glassey, A. Romera, and J. Jago ........................................................................................................ 191 2 SPATIAL AND TEMPORAL FREE-RANGING COW BEHAVIOUR PRE AND POST-WEANING. D. M. Anderson, C. Winters, M. Doniec, C. Detweiler, D. Rus, and B. Nolen ...................................................... 192 3 EFFECT OF SUPPLEMENT ENERGY LEVEL ON FATTY ACID PROFILES AND MEAT QUALITY OF STEERS FINISHED ON WINTER ANNUAL PASTURE. H. O. Patino and F. S. Medeiros .................. 193 4 USE OF N-ALKANES AND LONG CHAIN ALCOHOLS TO ESTIMATE FORAGE INTAKE AND DIET COMPOSITION OF CATTLE GRAZING TWO FORAGE SPECIES. H. T. Boland and G. Scaglia ................................................................................................................................... 194 5 PROTEIN AND CARBOHYDRATE DEGRADATION CHARACTERISTICS AND RATIOS OF ANTHOCYANIDIN-ACCUMULATING LC-ALFALFA AND ALFALFA SELECTED FOR A LOW INITIAL RATE OF DEGRADATION IN GRAZING CATTLE. A.. Jonker, M. Gruber, Y. Wang, and P. Yu ................................................................................................................................................................ 195 6 GROWING AND FATTENING CATTLE ON THE NORTHERN GREAT PLAINS WITH RUMINALLY-PROTECTED FLAXSEED. S. L. Kronberg, E. J. Murphy, R. J. Maddock3, and E. J. Scholljegerdes .......................................................................................................................................... 196 7 GRASS INTAKE OF GRAZING DAIRY COWS USING THE N-ALKANE TECHNIQUE. A. van den Pol-van Dasselaar and A. Hensen ........................................................................................................ 197 8 GRAZING IN EUROPE. A. van den Pol-van Dasselaar .............................................................................. 198 9 PATHWAY FOR THE ELIMINATION OF MELAMINE IN LACTATING DAIRY COWS. J. Shen, J. Wang, H. Wei, D. Bu, and P. Sun ....................................................................................................................... 199 10 COMPARE THE DIGESTIBILITY OF RUMEN UNDEGRADABLE PROTEIN OF CORN DDGS AND DDG IN INTESTINE OF CHINESE HOLSTEIN DAIRY COWS USING MOBILE NYLON BAG TECHNIQUE. Z. H. Yan, J. Q. Wang, D. P. Bu, H. Y. Wei, L. Y. Zhou, and P. Sun ........................................ 200 11 EFFECT OF SITE AND SOURCE OF LYSINE SUPPLEMENTATION ON NITROGEN METABOLISM IN BEEF CATTLE. Y. Q. Guo, Y. D. Zhang, J. Q. Wang, D. P. Bu, K. L. Liu, and T. Hu ................................................................................................................................................................ 201 12 SUPPLEMENTATION WITH PROTEIN AND VARIOUS FEED ADDITIVES IMPROVES IN SITU DEGRADATION CHARACTERISTICS OF BERMUDAGRASS AND BUFFELGRASS IN STEERS CONSUMING A LOW-QUALITY FORAGE DIET. K. C. McCuistion, B. D. Lambert, T. A. Wickersham, R. O. Dittmar, L. Wiley, and L. Dobson................................................................................................................. 202 13 CASE STUDY: MOLASSES AS THE PRIMARY ENERGY SUPPLEMENT ON AN ORGANIC GRAZING DAIRY FARM. K. Hoffman, L. E. Chase, and K. J. Soder ............................................................. 203 14 POTENTIAL OF LEGUMES AS SUBSTITUTES FOR NITROGEN FERTILIZER IN SUMMER STOCKER GRAZING SYSTEMS. R. R. Reuter, J. T. Biermacher, J. K. Rogers, T. J. Butler, M. K. Kering, J. R. Blanton, and J. A. Guretzky ........................................................................................................................... 204 15 EFFECTS OF CONVENTIONAL AND GRASS FED SYSTEM ON CARCASS TRAITS AND PERFORMANCE OF ANGUS STEERS. G. Cruz, G. Acetoze, and H. Rossow.............................................. 205 16 ASPECTS OF CARCASS COMPOSITION AND ORGAN WEIGHTS OF ANGUS STEERS FINISHED ON GRASS AND HIGH GRAIN DIETS. G. Acetoze, G. D. Cruz, and H. A. Rossow .................................... 206 17 PERFORMANCE AND EFFICIENCY ON PASTURE OF CATTLE WITH DIVERGENT PHENOTYPES FOR RESIDUAL FEED INTAKE. T. D. A. Forbes, A. D. Aguiar, L. O. Tedeschi, F. M. Rouquette, Jr., G. E. Carstens, and R. D. Randel .......................................................................................... 207 18 EFFECTS OF SAND SAGEBRUSH CONTROL ON STOCKER CATTLE PERFORMANCE ON A SOUTHERN MIXED PRAIRIE COMMUNITY. S. A. Gunter, T. L. Springer, E. Thacker, and R. L. Gillen ...................................................................................................................................................... 208 19 PASTURE MANAGEMENT EFFECTS ON NONPOINT SOURCE POLLUTION OF MIDWESTERN PASTURES. D. A. Bear, J. R. Russell, and D. G. Morrical .................................................... 209 20 SUPPLEMENTATION OF RUMINALLY UNDEGRADABLE PROTEIN TO MAINTAIN ESSENTIAL AMINO ACID SUPPLY DURING NUTRIENT RESTRICTION ALTERS CIRCULATING ESSENTIAL AMINO ACIDS OF BEEF COWS IN EARLY TO MID-GESTATION. A. M. Meyer, J. S. Caton, M. Du, and B. W. Hess........................................................................................................................ 210 21 EFFECT OF FORAGE TYPE ON PERFORMANCE OF WEANLING BEEF STEERS. G. Scaglia, B. Corl, W. S. Swecker, Jr., and A. Lillie ............................................................................................................... 211 22 ESTIMATES OF DRY MATTER INTAKE OF BEEF STEERS GRAZING HIGH QUALITY PASTURES USING ALKANES. G. Scaglia and H. Boland .............................................................................. 212 23 EFFECT OF ESTABLISHMENT METHOD OF WHEAT PASTURE AND FALL STOCKING RATE ON PERFORMANCE OF GROWING STEERS. P. Beck, M. Morgan, T. Hess, D. Hubbell, M. Anders, and B. Watkins ....................................................................................................................................................... 213 24 GROWTH PERFORMANCE EFFECTS OF MINERAL VS. SALT SUPPLEMENTATION ON STOCKER CATTLE GRAZING SPRING-SUMMER PASTURES OVER TWO CONSECUTIVE YEARS. P. J. Guiroy, S. E. Showers, and B. McMurry ........................................................................................................ 214 25 USE OF DRIED DISTILLERS GRAINS AS A SUPPLEMENTAL FEEDSTUFF FOR GROWING CATTLE ON WHEAT PASTURE. E. D. Sharman, P. A. Lancaster, G. W. Horn, and J. T. Edwards ............. 215 26 A GENETIC MARKER FOR RESISTANCE TO FESCUE TOXICOSIS IN BEEF CATTLE. C. J. Kojima, J. C. Waller, D. E. Spiers, R. L. Kallenbach, B. T. Campbell, T. A. Cooper, and J. K. Bryant ....... 216 27 PERFORMANCE OF CATTLE GRAZING SWITCHGRASS OR COMBINATIONS OF BIG BLUESTEM AND INDIANGRASS DURING THE SUMMER IN THE MID-SOUTH. W. M. Backus, J. C. Waller, P. D. Keyser, G. E. Bates, C. A. Harper, F. N. Schrick, and B. T. Campbell .................................... 217 28 BOTANICAL COMPOSITION, NUTRITIONAL VALUE, VOLUNTARY FEED INTAKE ESTIMATION AND BEHAVIORAL HABITS OF CATTLE GRAZING ON RANGE PASTURES. D. Rodriguez-Tenorio, R. Gutierrez-Luna, R. D. Valdez-Cepeda, F. G. Echavarria, M. A. Salas, J. I. Aguilera-Soto, M. A. Lopez-Carlos, C. F. Arechiga, and J. M. Silva-Ramos ................................................................................ 218 Author Index ......................................................................................................................................... 219 Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 ISSUES IN GRAZING LIVESTOCK NUTRITION T. DelCurtoŧ and K. C. Olson* ŧ Eastern Oregon Agricultural Research Center, Oregon State University, Union 97883; *West River Agricultural Research and Extension Center, South Dakota State University, Rapid City 57702 difficulty in maintaining esophageal fistulated cattle, increased scrutiny from Institutional Animal Use and Care Committees, and in turn, the relative ease of maintaining rumen cannulated cattle. One of the most fundamental equations for a grazing animal nutritionist is Forage Intake = Fecal Output /(1 – diet digestibility). This equation is simple; however, the components (fecal output and digestibility) required to make it biologically meaningful have plagued grazing livestock nutritionist for the past 60 years. The ability to estimate fecal output has obvious problems. The most direct technique is to use fecal bags and collect the feces either daily or twice daily depending on diet and size of the animal. This technique may work well for studies conducted with animals in confinement and studies with animals grazing small pastures or paddocks, but this approach has limited application to animals in extensive free-ranging environments. An alternative approach is to use markers, a topic reviewed in previous grazing conferences (Cochran et al., 1987; Pond et al., 1987). However, many of the marker techniques have limitations with animals in extensive free-ranging environments. Likewise, digestibility estimates can be challenging and the best technique often depends on the research needs and experimental conditions. Whenever the research questions involve the need for “free-ranging” herbivory, the techniques used to assess fecal output and (or) digestibility become more difficult. Nutritionists working with grazing animals have historically focused on chemical composition of diets because of the direct relationship to predicting and explaining animal performance. Often, little attention was devoted to botanical composition of diets. However, a growing number of researchers are interested in techniques to determine botanical composition of diets. In fact, grazing behavior research is a growing area of emphasis and is reflected in an increased number of papers with reference to botanical composition of diets. Holechek et al. (1982) have provided a review of techniques to estimate botanical composition of diet. Both bite count (Wickstrom et al., 1984; Canon et al., 1987) and microhistological techniques (Sparks and Malechek, 1968; Holechek et al., 1982) have useful applications with grazing animal research. With increasing demands to demonstrate that domestic livestock production is compatible with biodiversity of native vegetation and wildlife, it would seem logical that diet composition research will continue as long as this research is clearly connected to available vegetation. This research, in turn, will be used to develop preference indices that can INTRODUCTION Grazing livestock nutrition is an area of academic study that is challenging, deals with dynamic ecosystems, and, currently seems to function across academic disciplines. Although the US Land Grant system has demonstrated 150 years of success in promoting agriculture in the U.S. and the World, the study of grazing livestock nutrition has never received the focus and (or) funding that this discipline deserves. When evaluating cow/calf production and the relative time spent in grazing scenarios versus feeding conserved forages, most of a beef cow’s productive life is spent grazing improved pastures, native meadows, and rangelands. Most beef cattle managers recognize that at least half of a cow’s productive life is spent in free-range grazing scenarios and most low cost production systems have their mature cows grazing greater than 75% of the year. Escalating prices of fossil fuel coupled with increased costs of cereal grains and other feed commodities makes extending the grazing period and reducing cow costs ever more important. Finally, the future of ruminant livestock production may be a direct result of the ability to convert forages and other high-fiber food by-products to usable high-quality protein sources for human diets. Simply put, grazing livestock nutrition is going to be paramount to the success of the ruminant livestock industry in the future. HISTORY OF GRAZING LIVESTOCK NUTRITION Clanton and Raleigh (1987) discussed the “history” of grazing livestock nutrition in the proceedings of the first Grazing Livestock Nutrition Conference. Although numerous studies are evident over the past 100 years, most of the literature specific to grazing livestock has been published in the last 60 years. The earliest research appeared to focus on the quality of the available forage (e.g., Cook and Harris, 1950; Forbes and Garrigus, 1950). The ability to measure quality of consumed diets did not become viable until the development of esophageal (Torell, 1954) and ruminal (Lesperance et al., 1960) fistula and cannula sampling techniques. Both approaches have advantages and disadvantages, and it seems the use of rumen cannulated cattle has become more popular among researchers (Olson, 1991). This is likely due to the 1 Corresponding author: [email protected] 1 ideas among attendees before and after sessions. To promote attendance, they wanted to hold it immediately preceding the National American Society of Animal Science (ASAS) Meeting in a desirable, resort location near the ASAS site. Thus, the first Grazing Livestock Nutrition Conference (GLNC) was held in 1987 in Jackson, Wyoming, immediately before National ASAS hosted by Utah State University in Logan. The goal of the resort location was to provide an enjoyable atmosphere where attendees could interact in a relaxed setting, further promoting collegiality and communication. Another key element was the publication of proceedings that would provide state-of-the-art documentation on topics related to grazing livestock nutrition. The 1st GLNC was a great success. The proceedings have been an invaluable asset and have been heavily cited. From there, plans developed to hold recurring GLNCs every four years, always immediately before National ASAS when it was located in the Western Section of ASAS (back in those days, the national meeting rotated among the 4 sections in a regular fashion). Thus, the 2nd GLNC was held in 1991 in Steamboat Springs, Colorado before National ASAS was in Laramie, Wyoming, and the 3rd GLNC was held in 1996 near Custer, South Dakota before National ASAS was in Rapid City. An additional key element added in the 2 nd and 3rd Conferences was inclusion of International keynote speakers. This was intended to augment the original key element of fostering communication and collegiality to an international level. Themes of the 1st GLNC included: “Factors Influencing Energy Requirements of Livestock”, “Techniques for Estimating Nutrient Intake and Utilization”, “Factors Influencing Digesta Flow”, and “Supplementation Practices”. Those themes were (and still are) core to grazing livestock nutrition, i.e., they included core elements of ruminant nutrition and grazing management presented in the context of the grazing animal. The 1st GLNC provided the opportunity to gather and present state-of-the-art information from experts in each of these themes. For example, the section on “Factors Influencing Energy Requirements of Livestock” included 6 papers (Ferrell and Jenkins, 1987; Havstad and Doornbos, 1987; Morris and Sanchez, 1987; Adams, 1987; Robertshaw, 1987; Young, 1987) that covered major topics related to increases in energy requirements as a result of the grazing environment. A particularly meaningful theme was “Techniques for Estimating Nutrient Intake and Utilization”. The mid- to late-1980’s was a period of intense technique development in ruminant nutrition that was of particular interest to grazing livestock nutritionists. The papers in this section (Cochran et al., 1987; Pond et al., 1987; Goetsch and Owens, 1987) played key roles in establishing the state of development at the time of the conference, and stimulated further development beyond that point. The 2nd and 3rd GLNC built upon the solid foundation that the 1st GLNC provided. Some themes of these subsequent conferences were similar and built upon the 1st, while other themes were novel. For example, energetic requirements in the context of the grazing environment can be found in all three conference proceedings (i.e., previously cited papers in the 1st GLNC, Crooker et al., and will be used to model the long term effects of herbivores on plant diversity and successional trends. Some of the most interesting past research with grazing animals involved supplementation research in free-ranging conditions (Clanton and Raleigh, 1987). These projects demonstrated that we could design supplementation strategies that optimized the use of diverse native or improved rangelands with beef cattle and sheep production. These studies were well received by beef cattle managers and are still relevant to current production systems. Nutritionists should always be cognizant that the best indicator of diet quality is actual animal performance. Estimates of intake and chemical composition of diet are only tools to predict animal performance under varying range conditions and management strategies. It is our opinion that recent research has lost sight of the fact that the best data is actual performance data (weight change, body condition change, reproductive efficiency, etc.) conducted in “real world” management systems. Perhaps the publication demands and economic limitations of the modern Land Grant University have hindered our willingness to invest time and resources into more applied, performance oriented research. One of the more significant developments in grazing livestock nutrition research in North America relates to the establishment of regional projects and (or) coordinating committees. These efforts are designed to foster multiuniversity and multi-agency collaborations that more effectively address regional issues and needs. Range livestock nutrition was first covered by the establishment of the W-34 project in 1955 (Clanton and Raleigh, 1987). Over the past 55 years, this project has oscillated from regional project to coordinating committee status and currently is a multi-state research project, W-1012. This group has produced numerous projects and publications that addressed techniques to estimate intake, diet quality, and diet selection, as well as research that relate to range livestock management. Clearly, the grazing livestock nutrition discipline is still relatively new and has numerous areas that could be improved upon. For individuals that have dedicated a substantial portion of their lives to this discipline, you likely understand the great rewards and frustrations in working with grazing animal nutrition and management. For younger people possibly interested in grazing animal nutrition research, there are certainly opportunities to make substantial contributions to this discipline and, in the process, you may also provide a huge service to the range livestock industry. HISTORY OF THE GRAZING LIVESTOCK NUTRITION CONFERENCE In the early to mid 1980s, a group of grazing livestock nutritionists from throughout the western U.S. recognized the opportunity and need for a conference to discuss the state of the art and future opportunities in grazing livestock nutrition. That group envisioned several key elements for the meeting. They wanted it to last 2 to 3 days, first so that in-depth treatment could be provided for important topics, but also to promote communication and cross-fostering of 2 grazing setting (e.g., Brock and Owensby, 2000). Opportunities for rapid development will continue, and the evaluation of grazing livestock use of their environment and the environment’s influence on patterns of nutrient harvest by livestock will be important in future development of our knowledge in grazing livestock nutrition. 1991; Frisch and Vercoe, 1991; Hennessey, 1996; Caton and Dhuyvetter, 1996). A wide variety of presentations about supplementation practices are included in all three proceedings. In fact, the 3rd GLNC was entirely focused on supplementation. By following the progression through the three proceedings, a reader can follow the evolution of thinking on specific topics over that 9-yr time span. Particularly with the recurring supplementation theme, the development of knowledge of utilization of supplemental protein, energy, and minerals is evident. For example, presentations about protein supplementation progress from being primarily N and CP based by Petersen (1987) and Owens et al. (1991) to an MP basis in 1996 (Paterson et al., 1996). This new emphasis in 1996 coincided with the release of NRC (1996), which converted calculation of beef cattle protein requirements from a CP to an MP basis. In turn, influence of this knowledge on applied aspects such as development of supplementation practices is evident, particularly in the 3rd GLNC proceedings (e.g., Torell and Torell, 1996). Another development that is evident over the span of the 3 proceedings is the increased knowledge of grazing behavior and appreciation for its influence on nutrient intake by grazing ruminants (i.e., Dougherty, 1991; Provenza, 1991; Ruyle and Rice, 1996; Rittenhouse and Bailey, 1996). After the 3rd GLNC, no further conferences have been convened. Members of WERA-110, a regional education/extension and research committee (recently reapproved as a multi-state research project, W1012) comprised of several scientists focused on grazing livestock nutrition, recognized the need to re-establish the GLNC. The GLNC had fulfilled an important role in the past, and the committee envisioned that role continuing. Thus, planning began in 2007 to convene this 4 th GLNC in 2010. The intent has been to reinstitute the key elements of past GLNC, including the multi-day event held in a resort setting near to and immediately before commencement of the annual meeting of ASAS. Additionally, the International speaker element has also been retained. The intent is to rekindle the mission of the prior GLNC to address and discuss the current state and future of the discipline in a setting that will foster collegiality and communication. Rumen Microbial Genomics The various “omics” fields have been active in animal science, but have often been considered too basic for the applied nature of grazing livestock nutrition as a discipline. However, the ability to link the benchtop nature of these molecular disciplines with the applied disciplines is rapidly advancing. This provides an exciting opportunity, particularly in the realm of rapidly expanding the ability to evaluate rumen microbiology, and it is especially exciting because of the opportunity to bring this ability into the field setting. Much like some of the papers in the 1st GLNC prompted rapid development of better grazing nutrition techniques; the inclusion of this theme in the 4 th GLNC will hopefully foster linkage between these fields of study and a similar rapid development of techniques and new knowledge. New Knowledge about Metabolism from Molecular to Organism Levels Since the last GLNC, knowledge of animal metabolism and techniques to evaluate it have continued to develop. These developments range from whole organisms to tissues, cells, and molecules. The integration of techniques across this gradient has vastly improved our ability to understand the metabolic mechanisms that drive the empirical observations of the past. There is a need to revisit themes of past GLNC and consider those relationships with metabolic drivers of processes and responses in grazing livestock. ISSUES AND CHALLENGES IN GRAZING LIVESTOCK NUTRITION SINCE 1996 Grazing Intake Techniques EVOLUTION OF GRAZING LIVESTOCK NUTRITION SINCE 1996 Geospatial Methodologies Range science and other natural resource disciplines have made rapid advancement in recent years using geospatial tools such as graphic information systems (GIS) and global positioning systems (GPS) to evaluate spatial relationships across landscapes of rangelands. The ability to evaluate relationships among plant communities, soils, geologic features, and physiographic features has been enlightening (Johnson et al., 2000; Boyce et al., 2002). Further development with use of GPS to track animal 1. locations (both domestic livestock and wildlife) in relation to these landscape features has further added to the value of geospatial techniques in considering livestock in the 3 Unfortunately, techniques to estimate grazing animal intake have not improved over the past 15 years, and in many respects, have lost ground from a practical standpoint. One of the most versatile external markers, chromic oxide, has been listed as a potential carcinogen and, as a result, the purchase and use of this marker has become limited. Chromic oxide was a marker that was easily added to supplements and (or) rations and was easily detected with atomic absorption using an acetylene head. In addition, the manufacture of sustained release boluses (chromic oxide and alkane; Captec Ltd., Auckland, New Zealand) has been discontinued. The loss of manufacture of sustained release boluses is presumably due to one or more of the following factors: 1. Chromic oxide being listed as a potential carcinogen; 2. The inability of researchers to measure alkanes in a repeatable manner; and and quality (Putman and DelCurto, 2007). Successful beef producers are not the ones who wean the heaviest calves, or obtain 95% conception, or who provide the most optimal winter nutrition. Instead, successful producers are those who demonstrate economic viability and sustainability (Figure 1) while maintaining ecological and social values. Western US rangelands have historically been managed to accommodate livestock production. However, Congress has altered the framework that governs federal land management with the passage of the Multiple Use Act (1968), National Environmental Policy Act (1969), Clean Water Act (1972), and the Threatened and Endangered Species Act (1973). The continued use of public and private rangelands depends on our ability to develop sustainable systems that maintain or enhance biological diversity of forages, riparian function, and wildlife across the western US. Grazing livestock nutrition and management must develop systems for economic viability that also maintain biological diversity (vegetation and wildlife) and the traditions and integrity of the industry. Research that is grounded in economic and ecologic sustainability should be encouraged and supported. Recent reviews evaluating the management of livestock distribution and applied management strategies for optimal distribution on arid rangelands provide relevant background for this discussion (Bailey, 2005; DelCurto et al., 2005). One of our biggest challenges may be in defining better approaches to measure or interpret cow efficiency. While academia and other support industries are putting greater efforts into “cow efficiency,” we still struggle with the concept and, to date, refuse to accept that the ideal cow for one nutritional environment may not be ideally suited for another environment. Likewise, we need to provide a framework for our beef industry to make genetic selection decisions that optimizes production for specific nutritional environments. We still have a chronic problem with producers who choose genetics that exceed their environmental potential and subsequent reproductive efficiency or failure effectively limits their desired genetic improvement. Obviously, we have work to do before these goals and aspirations will be fully realized. 3. The low demand for sustained release boluses in respect to overhead and production cost. Additionally, the interest and time devoted to developing new techniques or refining existing techniques seems to be declining. This may be driven by an inability to find external funding for research technique development and (or) the decreased number of researchers devoted to this area of research. Inadequacy of NRC (1996) For many range livestock nutritionists, the 1996 Beef NRC has limited application to applied cow/calf nutrition. First, for grazing beef cattle in extensive scenarios where the cattle are adapted to limited nutritional environments, meeting the animal’s CP and energy requirements is the first priority. Balancing protein in terms of metabolizable protein, rumen undegradable protein (RUP), and rumen degradable protein (RDP) is beyond the needs of the average range livestock manager. Additionally, most grazing livestock nutrition issues deal with supplementation strategies that optimize the use of dormant, low-quality forage. In these scenarios, the intake formulas used in the 1996 Beef NRC do not accurately predict the forage intake response to protein supplementation. This may be due to the fact that microbial efficiency is a key component to these formulas and is based primarily on digestion studies conducted in the mid-west with low-quality warm season forages. Recently, Bohnert et al. (2007) demonstrated that the intake response to supplemental protein differed among warm season and cool season forage. This observation is presumably due to the differing cell wall structure despite the fact that both forages were similar in CP and NDF values. Likewise, research with frequency of protein supplementation and the success of less frequent strategies (Huston et al., 1999; Bohnert et al., 2002) suggest that protein fractions (RUP vs. RDP) are not necessarily important to mature cow nutrition in an extensive rangeland production environment. CHALLENGES TO GRAZING LIVESTOCK INDUSTRY FUNDING CHALLENGES FOR LAND GRANT UNIVERSITY RESEARCH Historically, grazing animal nutrition has focused on optimizing beef cattle performance with maximal use of rangeland forages. Beef production is optimized in that the goal usually involves minimizing supplemental inputs and harvested forages while maximizing the use of native and (or) introduced vegetation by the grazing animal. We have all heard the comments that we need to think of the grazing ruminant as the “harvester” of the available forage. The “ideal” or most efficient cow (or ewe) is the one that works best in the specific rangeland environment and requires the least amount of external nutritional input but still produces a viable, wholesome food product. Obviously, this creates difficulty in defining “what is the ideal animal” because the answer is specific to the nutritional environment that is provided by the rangeland environment and the management preference. Furthermore, the beef industry in the western U.S. is highly dependent upon variable arid environments and the subsequent effects on forage supply An unfortunate legacy since the last GLNC has been continual erosion of financial support of public universities in general, and the Land Grant University system throughout the U.S. has not been immune from this loss. This means there are fewer scientists using far fewer dollars to study the nutrition of grazing livestock now than there were before and during the era of the previous GLNC. This makes one pause about the sustainability of the field of study as well as the need and opportunity for further GLNC after the 4th. Grazing livestock nutritionists will need to recognize and embrace opportunities and challenges for the discipline to not only survive, but thrive. Novel approaches to garnering external funding and new partnerships with other disciplines will play a role in the future of grazing livestock nutrition as a viable discipline. 4 GOALS OF THE 4TH GRAZING LIVESTOCK NUTRITION CONFERENCE Cook, C. W. and L. E. Harris. 1950. The nutritive value of range forage as affected by vegetation type, site, and state of maturity. Utah Agr. Exp. Sta. Tech. Bull. 344. Crooker, B. A., P. T. Anderson, and R. D. Goodrich. 1991. Maintenance energy requirements and energetics of tissue deposition and mobilization in cattle. Pages 1-12 in Proc. 2nd Grazing Livestock Nutrition Conference. Oklahoma Agr. Exp. Stn. Pub. MP-133. DelCurto, T., M. Porath, C. T. Parsons, and J. A. Morrison. 2005. Management strategies for sustainable beef cattle grazing on forested rangelands in the Pacific northwest. Invited synthesis paper. Rangeland Ecology Manage. 58:119–127. Dougherty, C. T. 1991. Influence of ingestive behavior on nutrient intake of grazing livestock. Pages 74-82 in Proc. 2nd Grazing Livestock Nutrition Conference. Oklahoma Agr. Exp. Stn. Pub. MP-133. Ferrell, C., and T. G. Jenkins. 1987. Influence of biological types on energy requirements. Pages 1-7 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Forbes, R. M. and W. P. Garrigus. 1950. Some relationships between chemical composition, nutritive value, and intake of forages grazed by steers and wethers. J. Anim. Sci. 9:354-362. Frisch, J. E. and J. E. Vercoe. 1991. Factors affecting the utilization of nutrients by grazing beef cattle in Northern Australia. Pages 198-212 in Proc. 2nd Grazing Livestock Nutrition Conference. Oklahoma Agr. Exp. Stn. Pub. MP-133. Goetsch, A. L. and F. N. Owens. 1987. Impact of forage characteristics on microbial fermentation and ruminant performance. Pages 55-66 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Havstad, K. M. and D. E. Doornbos. 1987. Effect of biological type on grazing behavior and energy intake. Pages 9-15 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Hennessey, D. W. 1996. Appropriate supplementation strategies for enhancing production of grazing cattle in different environments. Proc. 3rd Grazing Livestock Nutrition Conference. M. B. Judkins and F. T. McCollum III (eds). Proc. West. Sec. Amer. Soc. Anim. Sci. 47(Suppl. 1):1-18. Holechek, J. L., M. Vavra, and R. D. Pieper. 1982. Botanical composition determination of range herbivore diets: A review. J. Range Manage. 42:248-251. Huston, J. E., H. Lippke, T. D. A. Forbes, J. W. Holloway, and R. V. Machen. 1999. Effects of supplemental feeding interval on adult cows in western Texas. J. Anim. Sci. 77:3057–3067. Johnson, B. K., J. W. Kern, M. J. Wisdom, S. L. Findholt, and J. G. Kie. 2000. Resource selection and spatial separation of mule deer and elk during spring. J. Wildl. Manage. 62:685-697. Lesperance, A. L., V. R. Bohman, and D. W. Marble. 1960. Development of techniques for evaluating grazed forage. J. Dairy Sci. 43:682-689. The first three conferences were tremendous successes that stimulated interest and continued collaboration for the discipline of grazing livestock nutrition. Simply put, it has been 14 years since the last “nutrition conference.” It is our sincere hope and intention that we will reinvigorate this conference and provide motivation and information to move our profession into the future. The organizing committee for the 4th Grazing Livestock Nutrition Conference has assembled an esteemed group of speakers to address the topics at hand. The charge to these speakers for their topic is to (1) provide historical perspective, (2) present the state-of-the-art knowledge, and (3) discuss future research needs and opportunities. LITERATURE CITED Adams, D. C. 1987. Influences of winter weather on range livestock. Pages 23-29 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Bailey, D. W. 2005. Identification and Creation of optimum habitat conditions for Livestock. Invited synthesis paper. Rangeland Ecology Manage. 58:109–118. Bohnert, D. W., C. S. Schauer, and T. DelCurto. 2002. Influence of rumen protein degradability and supplementation frequency on performance and nitrogen use in ruminants consuming low-quality forage: Cow performance and efficiency of nitrogen use in wethers. J. Anim. Sci. 80:1629-1637. Bohnert, D. W., T. DelCurto, A. A. Clark, M. L. Merrill, S. J. Falck, and D. L. Harmon. 2007. Protein supplementation of ruminants consuming low-quality cool or warm-season forage: Differences in intake and digestibility. Proc. West. Sec. Am. Soc. Anim. Sci. 58:217-220. Boyce, M. S., P. R. Vernier, S. E. Nielsen, and R. K. A. Schmiegelow. 2002. Evaluating resource selection functions. Ecological Modeling 157:281-300. Brock, B. L. and C. E. Owensby. 2000. Predictive models for grazing distribution: A GIS approach J. Range Manage. 53:39-46. Canon, S. K., P. J. Urness, and N. V. DeBule. 1987. Foraging behavior and dietary nutrition of elk in burned aspen forest. J. Range Manage. 40:433-438. Caton, J. S., and D. V. Dhuyvetter. 1996. Manipulation of maintenance requirements with supplementation. Proc. 3rd Grazing Livestock Nutrition Conference. M. B. Judkins and F. T. McCollum III (eds). Proc. West. Sec. Amer. Soc. Anim. Sci. 47(Suppl. 1):72-82. Clanton, D. C. and R. J. Raleigh. 1987. Forty years of grazing livestock nutrition research. Pages 151-156 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Cochran, R. C., E. S. Vanzant, K. A. Jacques, M. L. Galyean, D. C. Adams, and J. D. Wallace. 1987. Internal markers. Pages 39-48 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. 5 Morris, J. G. and M. D. Sanchez. 1987. Energy expenditure of grazing ruminants. Pages 17-22 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. NRC. 1996. Nutrient Requirements of Beef Cattle. 7 th rev. ed. Natl. Acad. Press, Washington, DC. Olson, K. C. 1991. Diet sample collection by esophageal fistula and rumen evacuation techniques. J. Range Manage. 44:515-519. Owens, F. N., J. Garza, and P. Dubeski. 1991. Advances in amino acid and N nutrition in grazing ruminants. Pages 109-137 in Proc. 2nd Grazing Livestock Nutrition Conference. Oklahoma Agr. Exp. Stn. Pub. MP-133. Paterson, J., R. Cochran, and T. Klopfenstein. 1996. Degradable and undegradable protein responses of cattle consuming forage-based diets. Proc. 3rd Grazing Livestock Nutrition Conference. M. B. Judkins and F. T. McCollum III (eds). Proc. West. Sec. Amer. Soc. Anim. Sci. 47(Suppl. 1):94-103 Petersen, M. K. 1987. Nitrogen supplementation of grazing livestock. Pages 115-121 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Pond, K. R., J. C. Burns, and D. S. Fisher. 1987. External markers - Use and methodology in grazing studies. Pages 49-53 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Provenza, F. D. 1991. Behavior and nutrition are complementary endeavors. Pages 157-169 in Proc. 2nd Grazing Livestock Nutrition Conference. Oklahoma Agr. Exp. Stn. Pub. MP-133. Putman, D. H. and T. DelCurto. 2007. Forage systems for Arid Zones. Pages 323-339 In: R. F. Barnes, C. J. Nelson, K. J. Moore, and M. Collins (eds). Forages, Volume II. The Science of Grassland Agriculture 6th Edition. Iowa State Press. Rittenhouse, L. R. and D. W. Bailey. 1996. Spatial and temporal distribution of nutrients: Adaptive significance to free-grazing herbivores. Proc. 3rd Grazing Livestock Nutrition Conference. M. B. Judkins and F. T. McCollum III (eds). Proc. West. Sec. Amer. Soc. Anim. Sci. 47(Suppl. 1):51-61. Robertshaw, D. 1987. Heat stress. Pages 31-35 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. Ruyle, G. B. and R. W. Rice. 1996. Aspects of forage availability and short-term intake influencing range livestock production. Proc. 3rd Grazing Livestock Nutrition Conference. M. B. Judkins and F. T. McCollum III (eds). Proc. West. Sec. Amer. Soc. Anim. Sci. 47(Suppl. 1):40-50. Sparks, D. R. and J. C. Malechek. 1968. Estimating percentage dry weight in diets using a microscopic technique. J Range Manage. 21:264-265. Torell, D. T. 1954. An esophageal fistula for animal nutrition studies. J. Anim. Sci. 13:878-884. Torell, L. A. and R. C. Torell. 1996. Evaluating economics of supplementation practices. Proc. 3rd Grazing Livestock Nutrition Conference. M. B. Judkins and F. T. McCollum III (eds). Proc. West. Sec. Amer. Soc. Anim. Sci. 47(Suppl. 1):62-71. Wickstrom, M. L., C. T. Robbins, T. A. Hanley, D. E. Spalinger, and S. M. Parish. 1984. Food intakes and foraging energetics of elk and mule deer. J. Wildl. Manage. 48:1285-1301. Young, B.A. 1987. Thermal factors influencing energy requirements of livestock. Page 37 in Proc. Grazing Livestock Nutrition Conference, Jackson, WY. Univ. Wyoming, Laramie. 6 Figure 1. Grazing Animal Nutrition needs to focus on research applications that provide balanced economic return, ecological integrity of rangelands, and social acceptability to beef cattle producers and land managers. 7 ISSUES IN GRAZING LIVESTOCK NUTRITION T. DelCurto and K. C. Olson Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 EMERGING METHODS IN RUMEN MICROBIOLOGY J. M. Brulcŧ, C. J. Yeomanŧ, K. E. Nelson*, and B. A. White1ŧ ŧ Department of Animal Sciences and the Institute for Genomic Biology, University of Illinois, Urbana 61801; * The J. Craig Venter Institute, Rockville, MD 20850 Estimates are that the rumen protozoa are present at a concentration of 1 x 106 per mL of rumen content. Rumen protozoa have been shown to impact the pH of the rumen (Veira, 1986) and may be responsible for up to 30% of ruminal fiber degradation (Demeyer, 1981). Rumen protozoal species have also been implicated in the removal and metabolism of toxic compounds that might present in the rumen (Veira, 1986). Obligately anaerobic fungi most closely related to the Chytridiomycete fungi of the order Spizellomycetales (Bowman et al., 1992) have been found in the rumen (zoospore population densities of 10 3 to 105 per mL divided into 4 genera). The anaerobic rumen fungi exhibit two basic forms: “polycentric” species in which nuclei are located throughout the hyphal mass, and “monocentric” species with a concentration of nuclear material in a zoosporangium (Brookman et al., 2000). The rumen anaerobic fungi are probably best studied in relation to their plant cell wall degrading enzymes or glycoside hydrolases (GH), which include an assortment of cellulases, xylanases, mannanases, glucosidases, and other glucanases. Recent studies have shown there is a strong similarity between the GH of rumen fungi and bacteria, suggesting that many of the GH of anaerobic fungi have been acquired by horizontal gene transfer from coresident bacteria (Garcia-Vallve et al., 2000). The rumen anaerobic fungi also produce several forms of esterases, which release esterified acetic and hydroxycinnamic acids from plant cell walls (Borneman et al., 1992; Fillingham et al., 1999). Archaea are estimated to comprise approximately 0.3 to 3% of the biomass in the rumen. To date only strictly anaerobic methanogens have been identified (Lin et al., 1997, Ziemer et al., 2000). Cultivation-based approaches have revealed just seven genera that include Methanobacterium, Methanobrevibacter, Methanomicrobium, Methanoculleus and Methanosarcina spp. although a greater diversity is expected (Janssen and Kirs, 2008). ABSTRACT: Advances in genomics and next-generation sequencing technologies have spawned the field of metagenomics. Metagenomics enables DNA sequencing of all the microbial species (the microbiome) that are present in an environment using culture-independent approaches. Initial studies have revealed an unanticipated level of microbial diversity in soils and the oceans. These same levels of complexity are now being revealed in studies of the microbiomes from human gastrointestinal tracts (GIT), and the GIT of production animals, including poultry and cattle. Continuing cost-reductions associated with sequencing using these next-generation technologies makes it possible to query these microbiomes at an unprecedented depth of sequence coverage. Although the ruminant GIT microbiome has been examined using culture-based and 16S rRNA phylogenetic analyses, metagenomics represents an emerging method for advancing the field. Metagenomics, coupled with the generation of reference genome sequences of the predominant rumen microbes, holds great promise for increasing our understanding of the diversity of species present in the rumen of both dairy and beef production animals, and how these species can be manipulated to improve feed efficiency and reduce the associated environmental impacts, including mitigation of green house gases. Key words: metagenomics, rumen, genomics, DNA sequencing, microbiome INTRODUCTION The rumen microbial community, or microbiome, is characterized by its high population density, broad diversity and complexity of interactions. The rumen microbiome resides in what can be considered a fermentation vat that maintains a near constant environment with respect to pH, nutrient availability, and temperature. This microbiome has evolved the ability to degrade complex plant polymers and provide the animal (host) with VFA for growth. Additional products of this fermentation are H2, CO2, and CH4. Although bacteria are numerically predominant (10 11 cells per mL), the rumen also contains microbial representatives from the archaea, protozoa, and fungi (Hespell et al., 1997). 1 EMERGING PARADIGMS FOR STUDYING GASTROINTESTINAL TRACT MICROBIOMES It is estimated that the GIT microbiome of mammals contains approximately 1014 total bacteria and is composed of between 500 to 1000 species, although few of these have been characterized (Turnbaugh et al., 2007; Ley et al., 2008a,b). Most of these microbes have not been grown in culture, their identities remain unknown, and it is unclear which, if any, are involved in growth promotion or enhancement of host physiology. Furthermore, each bacterial species can be present at vastly different Corresponding author: [email protected] 10 fundamentally important that these challenges are overcome as microbial genome projects move forward. A list of microbial genomes that have been or are being sequenced from the rumen is shown in Table 1. Among the first of the rumen microbial genomes to be sequenced was the 2.3 Mbp genome of Mannheimia succiniciproducens (Hong et al., 2004). M. succiniciproducens MBEL55E is a non-motile, non-sporeforming, mesophilic and capnophilic gram-negative organism with a rod morphology (Lee et al., 2002). This bacterium efficiently fixes CO2 and produces large quantities of succinic acid (Lee et al., 2002). Actinobacillus succinogenes 130Z from the bovine rumen is also capnophilic, producing large amounts of succinate. A. succinogenes is also osmotolerant, and uses a range of sugars (Guettler et al., 1999), its 2,319,663 bp genome is available at the Joint Genome Institute website: (http://www.ncbi.nlm.nih.gov/sites/entrez?db=genome&cm d=Retrieve&dopt=Overview&list_uids=21212; http://genome.jgi-psf.org/actsu/actsu.home.html). Butyrivibrio proteoclasticus B316, formerly designated Clostridium proteoclasticum (Moon et al., 2008) is a proteolytic, hemicellulose-degrading bacterium that has been shown to play a significant role in the biohydrogenation of fatty acids (Wallace et al., 2006). B. proteoclasticus was isolated from the rumen contents of cattle grazing fresh forage in New Zealand (Attwood et al., 1996). It is has being sequenced to completion by AgResearch, New Zealand and is due for release this year. The completed Wolinella succinogenes genome sequence and its analysis was released in 2003 (Baar et al., 2003). Despite a close phylogenetic relationship to the GIT pathogens Helicobacter pylori and Campylobacter jejuni, it has not been shown to cause any diseases and is as such considered to be non-pathogenic to the host (Baar et al., 2003). The rumen fungi act synergistically with rumen bacteria and protozoa in the degradation of plant material. Nicholson et al. (2005) initiated genome sequencing and analyses of the rumen fungus Orpinomyces. Along with Piromyces (another rumen fungi), genome sequencing has proved difficult for rumen fungi due to their extraordinarily high proportion of AT residues relative to GC (~80 to 85 % of the total DNA and > 95 % in certain loci). Renewed attempts to capture the genomes of both Orpinomyces and Piromyces using next-generation sequencing methods are in progress through collaborations between the Joint Genome Institute (JGI) and the Pacific Northwest National Laboratory. The North American Consortium for Genomics of Fibrolytic Ruminal Bacteria (NACGFRB) was created in 2000 with support from USDA Initiative for Future Agriculture and Food Systems. Initial whole genome sequencing projects have been completed by the consortium for a number of physiologically important isolates from the rumen. The consortium has enhanced our genetic understanding of closely related rumen isolates through several suppressive subtractive hybridization (SSH) analyses (Galbraith et al., 2004, Qi et al., 2008; Morrison et al., 2009). The consortium has produced a closed genomes for Fibrobacter succinogenes strain S85 (Morrison et al., unpublished data), Prevotella ruminicola concentrations. For example, microbes such as the Bacterioides sp., are present at much higher concentrations (~1010 to 1012 per gram of feces) compared with many other microbes such as Escherichia coli (~104 to 105 per gram of feces). Because of the proportional differences in biomass and therefore overall metabolic contributions to the GIT, it has been long believed that the more highly represented bacteria (core microbiome) are more substantial contributors to nutrient utilization, animal health, and are of greater importance to food safety (Turnbaugh et al., 2007, 2009). Evolutionary relationships between the host and environment, including microbiomes, drive the composition of core (high abundance) and variable (low abundance) microbiomes. Core microbiomes are hypothesized to be composed of a high density, but relatively low diversity of microbial taxa, whereas the variable microbiome is composed of low density, but highly diverse microbial taxa (Turnbaugh et al., 2007, 2009). This has lead to the NIH-funded Human Microbiome Project (HMP; http://nihroadmap.nih.gov/hmp/) whose primary goals are to determine if there is a core human microbiome (Turnbaugh et al., 2007). In addition to this, the aims are to understand the changes in the human microbiome that can be correlated with human health, and to develop new technological and bioinformatics tools to support these goals. Initial sequencing objectives for this program are in place, and include the development of a microbial genome reference set comprising at least 600 to 1,000 genomes. Additionally, 16S rDNA phylogenetic analyses followed by metagenomic sequencing are being used to characterize the microbial communities from the different body sites of a pool of 250 individuals. It is this emerging paradigm for analyzing microbiomes that is becoming the model for studying other GIT microbiomes like the rumen. Therefore, any future large-scale comparative analyses of microbial community structure and functional metabolic potential of the rumen microbiome will rely on, and be greatly advanced by the recent technological advances and powerful genomic and metagenomic approaches outlined in the HMP. To fully explore the microbial complexity and environmental impact of such a diverse biome, a robust experimental approach would utilize a combination of sequencing technologies; 1) generation of reference microbial genome sequences; 2) 16S rDNA phylogenetic analysis; and 3) metagenomic sequencing of select samples. REFERENCE GENOMES FROM THE RUMEN MICROBIOME Phylogenetic and metabolic sequence information still relies heavily on reliable curation of complete genomes in the databases. Currently, the National Center for Biotechnology Information (NCBI) has listed 1,094 complete microbial genomes and 2,595 microbial genomes in progress. Sequencing technologies are becoming more affordable and with the increasing read-lengths from pyrosequencing, storage of such massive data sets and ensuring data curation is already becoming an issue. It is 11 approaches oriented at specific subpopulations have focused on targeting specific phyla and gene families. A SSU rDNA survey specifically oriented toward the bovine rumen sequenced 84 full-length clones generated from the rumen contents of Holstein cows (Tajima et al., 1999). Comparing these sequences to the ribosomal database project (RDP) resulted in 56% of the generated sequences showing less than 90% similarity to known species, highlighting an untouched diverse population yet to be sampled. Tajima et al. (2000) also found a shift in the overall rumen bacterial communities coincident with a shift from a primarily hay-based diet to a grain diet. In total, 150 full-length SSU rDNA sequences were generated in this study, revealing that in both diets, low G+C gram-positive bacteria were dominant, representing 90.2% and 72.4% of the 16S rDNA population, respectively. In the hay diet some low G+C gram-positive bacterial sequences clustered with known cellulose-degraders Ruminococcus flavefaciens and Ruminococcus albus. Surveys using SSU rDNA have emphasized the vastly and diverse microbial component that has yet to be cultured (Whitford et al., 1998; Kocherginskaya et al., 2001; Koike et al., 2003; Edwards et al., 2004). A comparative analysis of all published rumen libraries found that only 11% of the sequences found within these studies possessed similarity to cultured members (Edwards et al., 2004). However, in phylogenetic analysis, it is apparent that the rumen community appears to be dominated by low G+C grampositive bacteria and the Cytophaga-FlavobacteriumBacteriodes (CFB) group. In one of the most comprehensive SSU rDNA surveys to date, Yu et al. (2006) created a ribosomal sequence tag library (a SSU rDNA library composed of the concatemers of hypervariable region 1) increasing the amount of phylotypes sampled and sequence information retrieved (Yu et al., 2006). The 190 clones sequenced resulted in 1055 ribosomal sequence tags (RSTs) and 236 unique phylotypes with 95% sequence identity. Rarefaction analysis estimated the ruminal diversity to comprise no greater than 353 different phylotypes at 95% similarity, similar to previous estimates (341 OTUs; Edwards et al., 2004) indicating less than 25% coverage of bacterial diversity. An additional analysis of the rumen archaeal population presence by SSU rDNA libraries resulted in the generation of 44 clones (Tajima et al., 2001b). Although rumen archaea appear to be dominated by Methanobrevibacter and Methanosarcina species, the SSU rDNA analysis resulted in sequences that formed a novel cluster of archaea that did not associate with these methanogens (Janssen and Kirs, 2008). To specifically quantify overall populations of microbes or specific subgroups, competitive or real-time PCR has been employed. Competitive PCR enumerated the three major cellulolytic bacteria in the rumen (Fibrobacter succinogenes, Ruminococcus flavefaciens and Ruminococcus albus) finding that F. succinogenes is present in greater concentrations (107 cells per mL) than Ruminococcus species (104-6) cells per mL (Koike and Kobayashi, 2001). Real-time PCR has also been used to monitor not only fluctuations in bacterial cellulolytic groups over time, but also fungal populations (Denman and McSweeney, 2006). strain 23 (Purushe et al., unpublished data) and draft (7-8x coverage) for R. albus strain 8 and Prevotella bryantii strain B14. The consortium also developed the FibRumBa database (http://www.jcvi.org/rumenomics/) as a resource designed for individuals with an interest in the microbial ecology of the rumen. The FibRumBa database provides an near-complete list of rumen microbial genome projects at the J. Craig Venter Institute (JCVI) that are currently underway or completed (Morrison et al., 2009). The site keeps a running list of the rumen genomes that have been completed to date by the JCVI and welcomes suggestions for improvement from the rumen community (Morrison et al., 2009). Most recently, the genome of R. flavefaciens FD-1 has been sequenced to 29x-coverage using both Sanger and 454 pyrosequencing and the draft sequence was assembled into 119 contigs providing 4,576,399 bp of unique sequence. The genome has a GC content of 45%, and is predicted to encode at least 4,339 ORFs with an average gene length of 918 bp. R. flavefaciens contains a cellulosome with at least 225 dockerin-containing ORFs, including previously characterized cohesin-containing scaffoldin molecules (Zhang and Flint, 1992; Flint et al., 1993; Aurilia et al., 2000; Jindou et al., 2006; Rincon et al., 2005). These dockerin-containing ORFs encode a variety of catalytic modules including glycoside hydrolases, polysaccharide lyases, and carbohydrate esterases. The genomic evidence indicates that R. flavefaciens FD-1 has the largest known number of fiber-degrading enzymes likely to be arranged in a cellulosome architecture (Berg Miller et al., 2009). PHYLOGENETIC STUDIES OF THE RUMEN MICROBIOME Traditionally, culture-based methods have been used to numerically and phenotypically classify organisms in the rumen microbiome. However, since such a low portion of the microbial component can be cultivated, this approach imposes severe limitations on studying complex microbial ecosystems (Hugenholtz and Pace, 1996). Moreover, the lack of a clear definition of a microbial species in addition to the role horizontal gene transfer plays in microbial biomes, renders these traditional taxonomic classifications no longer applicable (Hugenholtz and Pace, 1996; Hugenholtz et al., 1998). These difficulties have led scientists to conceptually define species based on a more robust biological organization using small subunit (SSU) rRNA sequence-based phylogenetic tree classification proposed by Carl Woese (Woese et al., 1990; Pace, 1997). Since the discovery and classification of the third domain of life (Fox et al., 1977; Woese and Fox, 1977), the SSU rDNA gene has not only been a useful tool for phylogenetic analysis but has also provided a framework for evaluating microbial communities without the need for prior cultivation and is further applied with more recent sequence-based technological advances. Using this technology, studies have provided significant insight into a variety of microbial ecosystems (Amann et al., 1995; Pace, 1997; Denman and McSweeney, 2006). Approaches employed in analyzing the overall rumen community structure have mainly used SSU rDNA surveys, while 12 In a diet-directed approach, real time-PCR was used to quantify 13 species of rumen of bacteria during the transition from a hay to grain diet (Tajima et al., 2001a). Those authors found F. succinogenes dominates in hay samples but was 57-fold reduced when the diet was switched to grain. Additionally, R. flavefaciens decreased 10-fold across diet transitions. In contrast, Streptococcus bovis, Selenomonas ruminantium, Prevotella ruminicola, and Prevotella bryantii increased in number by 7 to 375fold during the same diet transition. They also found differential amplification with some rumen bacteria suggesting that although some species predominate, they may not be found or accurately enumerated in SSU rDNA surveys. Although the use of oligonucleotide probes to observe bacteria populations is not a new technique (Stahl et al., 1988), taxa specific probes can offer a range of specificity across varying taxonomic levels. This targeted approach was taken in the generation of genus specific Fibrobacter SSU rRNA oligonucleotide probes (Lin et al., 1994; Lin and Stahl, 1995). Also, relative abundance measurements obtained from Ruminococci-specific probes were compared to numerical estimates using more traditional enumeration techniques (i.e., direct microscopic counts) for validation (Krause et al., 1999). In a more broad representation of all domains of life in ruminant animals, Lin et al. (1997) designed oligonucleotide probes to reveal that bacteria represented anywhere from 60% to 90%, Eukarya, 3% to 30% and Archaeal 0.5% to 3% of the rRNA populations (Lin et al., 1997). Targeted phyla approaches have utilized fluorescence in situ hybridization (FISH) to observe cellulolytic bacteria. Shinkai et al. (2007) found F. succinogenes adherent to hay stems while R. flavefaciens was found adherent to the surfaces of leaves, thereby indicating species specificity or preference in location adherence (Shinkai and Kobayashi, 2007). Since F. succinogenes is found in greater numbers in the rumen and possesses the ability to solubilize crystalline cellulose, a more recalcitrant form of cellulose, it is hypothesized that this organism makes a greater contribution to fiber degradation. However, highlighted in these studies is an unprecedented amount of rumen microbial diversity often missed due to probe specificity. meaning a full plate run may yield 0.5 Gb of high quality sequence. Thus, this technology has been adapted to allow deep SSU rRNA hypervariable tag sequencing (Dethlefsen et al., 2008; Huse et al., 2008) for use in analyses of vertebrate GIT systems. One reason is this approach yields very high information content (see for example, Neufeld et al., 2004; Kysela et al., 2005; Neufeld and Mohn, 2005; Sogin et al., 2006; Roesch et al., 2007). Pyrosequencing targeting hypervariable regions of the 16S rDNA allows for deeper surveys. This approach was recently validated for the human vertebrate GIT in studies of both lean and obese twins (Turnbaugh et al., 2009) and patients undergoing antibiotic treatment (Dethlefsen et al., 2008; Huse et al., 2008). METAGENOMIC APPROACHES Natural environments are composed of a genomically diverse set of organisms. Often these organisms, particularly those that are underrepresented, are missed in culture despite their potential to supply significant and (or) unique metabolic contributions in these complex environments (Ley et al., 2006; Sogin et al., 2006; Turnbaugh et al., 2006). Microbiologists began using culture-independent methods, metagenomics, to circumvent the low proportion (often less than 1%) of cultivatable members of a microbiome (Handelsman, 2004; Schloss and Handelsman, 2008). Metagenomics allows genomic access to the entire population of microorganisms and allows for independent analysis of these microbes in conjunction with their natural habitat. Traditional metagenomic analyses generally began with total extracted genomic DNA of that community. The DNA can be restriction digested, ligated into a vector and propagated in a host, often E. coli. For a sequence-driven analysis, clones can be chosen at random and subsequently sequenced. For functional-driven analyses, the clones can be screened for phylogenetic markers, enzymatic activity, or antibody binding. Heterologous gene expression then allows for physiological identification of small molecules or proteins. More recently, sequenced-based metagenomic approaches have strayed from cloning techniques, which introduce their own levels of bias, to a more random sequencing strategy, pyrosequencing (Ronaghi et al., 1996, 1998; Margulies et al., 2005). Regardless of the approach taken, metagenomics appears to be a very useful tool to answer some of the basic questions that have persisted since the 19th century: who’s there, how many of them are there, and what are they doing? More specifically, microbiologists and microbial ecologists want to go beyond single organism genomics and delve into a more encompassing view that not only includes the types and numbers of these organisms present, but also their function and (or) influence as a community in the environment. The early world of metagenomics began with traditional Sanger sequencing. As microscopic enumeration and colony counts were compared to the resulting numbers of microbes cultured, it became apparent that there was a large majority of organisms that were overlooked in these traditional studies (Handelsman, 2004; Schloss and Handelsman, 2007). A demand for genomic tools arose that DEEP SMALL SUBUNIT rRNA HYPERVARIABLE TAG SEQUENCING New techniques are constantly being developed to further elucidate the diversity of microbial communities. Pyrosequencing is a promising new technology developed by 454 Life Sciences (http://www.454.com/; Hyman, 1988; Ronaghi et al., 1996, 1998; Margulies et al., 2005). The 454 technology presents a major advantage over traditional sequencing technologies by eliminating the need for cloning vectors, which are known to be biased in terms of the impaired ability to clone certain DNA fragments (http://www.454.com/). This sequencing technology also reads through secondary structures readily, and has a capacity to produce very large amounts of sequence. Current estimates are that this instrument produces about 1.25 million sequence reads, with read lengths of ~ 400 bp, 13 combined results lend EGT data to serve as a “fingerprint” for each microbiome. The use of Sanger sequencing poses some issues with metagenomic analyses. Often, the ability to generate large amounts of sequence information was limited by the cost of sequencing ($8,000 to $10,000 per Mbp; Goldberg et al., 2006) and the need for cloning introduces a bias, resulting in the more abundant members of the community present more often in the number of reads compared with less abundant members. Additionally, cloning impedes the analysis of viral metagenomes since expression of virulence genes often kill the host (Edwards and Rohwer, 2005). Next generation sequencing technologies, like pyrosequencing, remove the need for cloning. This and the Solexa system are the landmark technologies that have completely changed the face of metagenomics. Using the Roche Titanium system, one can generate up to 1,250,000 reads per run, with >400 bp/read (0.5 Gbases of total sequence) for less than $10,000. While the Illumina GAII system produces shorter reads (<100 bp), it can result in over 50 GB/run. This system was recently applied to human GIT metagenomics by the MetaHIT group with very promising results (Qin et al., 2010). Its replacement, the HiSeq2000 will generate upwards of 100 bp fragments with up to to 200 Gb in a single run for less than $10,000. Edwards et al. (2006) were the first to apply random sample pyrosequencing to total environmental DNA collected from a subterranean environment, the Soudan Mine. Following a whole genome PCR amplification step, phylogeny was analyzed via SSU rDNA distributions and coding sequences were categorized into functional subsystems. A SSU rDNA clone library was generated for comparative analysis against the more traditional Sanger sequencing method. They generated approximately 350,000 sequences (average read length ~100 bp) for two samples that were in close proximity (100 m apart) to one another (Edwards et al., 2006). From these sequences, a high congruency was found between SSU rDNA sequences obtained from pyrosequencing and the taxonomic profile observed the SSU clone library. Additionally, the metabolic potential of these communities was determined via comparative analysis in the SEED database (Overbeek et al., 2005). The SEED database uses an approach based on subsystems, a set of functional roles often coordinating groups of genes found in a metabolic pathways or cell structure components. These subsystems are curated from complete genomes by an expert in that particular field. The subsystem analysis indicated that these two distinct communities possessed extremely different metabolisms. For example, oxidized and reduced samples were greatly overrepresented in corresponding subsystems (RodriguezBrito et al., 2006) and these metabolisms correlated with sampled environmental factors enabling the formation of a working microbiome model on how these microbial communities coexist. Similar to the Sanger sequencing study of Tringe et al. (2005), a functional metabolic meta-analysis of multiple metagenome data sets (over 14 million sequences from 45 microbiomes and 42 viromes), that differ significantly in biome origin, revealed differential representation of metabolisms based on environment (Dinsdale et al., 2008). would allow for a more accurate picture of the phylogenetic distribution of the microbial diversity present. Moving beyond SSU rDNA surveys, Sanger sequencing of large insert libraries, traditionally using single organism genomics, was applied to total community DNA (Rondon et al., 2000). In this study, heterologous gene expression using low copy bacterial artificial chromosomes (BAC) to sequence community DNA from a soil microbiome (Rondon et al., 2000) resulted in more than 1 gigabase pairs of DNA between two libraries. In addition to successfully finding sequences from a diverse array of taxa, they were able to screen the clones for functional diversity resulting in novel gene discovery. This enabled a tied study of genetics to functional expression for each of the selected clones. Another landmark study, the acid mine drainage (AMD) study demonstrated the ability to reconstruct genomes from a simple microbial community (Tyson et al., 2004). A biofilm community was found growing atop a water surface composed of H2SO4 and FeOH3, which the microbes had produced in response to an anthropogenic disturbance caused by mining. These microbes performing the reaction subsequently lowered the pH of the environment, limiting the number of organisms that coinhabit that location. This study generated a total of 76.2 megabase pairs (~100,000 reads) with an average insert size of 3.2 kb. The 384 SSU rDNA clones end-sequenced resulted in 3 bacterial lineages and 3 archaeal lineages. From about 85% of the shotgun reads assembled into scaffolds greater than 2kb, Tyson et al. (2004) discovered almost complete genomes for an iron-oxidizing bacteria and an acidophilic archaea, thereby proving that multiple genomes could be reconstructed from a simple community using metagenomic shotgun sequencing methods. Additionally, with the ability to recreate and compare contigs, Tyson et al. (2004) found evidence of horizontal gene transfer, determined potential proteins that indicated the physiological source of organismal chemical product generation, and found an underrepresented taxa group that may be the keystone for AMD community survival. More complex communities, such as the Sargasso Sea (Venter et al., 2004), the human metagenome (Gill et al., 2006; Turnbaugh et al., 2006, 2007; Kurokawa et al., 2007) and the termite hindgut (Warnecke et al., 2007), have been subjected to a shotgun sequencing approach. Although full microbial genomes are rarely reconstructed from these type of sequencing projects (generally less than 1% of reads overlap), much can be elucidated about their physiological contribution (Tringe et al., 2005). Tringe et al. (2005) performed a comparative analysis of multiple metagenomes using “predicted genes on small DNA fragments,” or environmental gene tags (EGT), to observe if gene sets for specialized functions would vary in number and type according to the environment sampled. Results indicated that each microbiome possesses gene sets similar to another microbiome if the metabolic demands are similar. By the same token, unique gene sets, such as bacteriorhodopsin found in Sargasso Sea samples and not in the soil metagenome, are restricted to select environments in which their gene products would be most utilized. These 14 pyrosequence data that had been derived from the fiberadherent and pooled liquid populations from 3 animals that had been fed the same diet. The samples had been obtained from 3 fistulated 5-yr old Angus Simmental Cross steers that had been maintained at the Illinois State University. Three liters of rumen contents were collected 1 h after feeding, and samples divided into Liquid and Fiber adherent. After DNA extraction, and quality control checks, the 4 total samples were sequenced using an early version of the 454 Life Sciences Genome Sequencer, GS20 (454 Life Sciences). Data that were obtained post sequencing were analyzed with the SEED annotation engine (http://metagenomics.nmpdr.org). For the assignment of carbohydrase enzymes, BLAST-based approaches against the CAZy database (www.cazy.org) were used. In parallel, 16S rRNA gene libraries were constructed with the universal prokaryotic primers 8FPL (AGTTTGATCCTGGCTCAG) and 1492RPL (GGYTACCTTGTTACGACTT). Matches were identified with BLASTN and rarefaction analysis was performed with DOTUR (http://www.plantpath.wisc.edu/fac/joh/dotur.html). These 16S rRNA gene sequences were compared against those present in the metagenomic libraries with Nonmetric Multidimensional Scaling (NM-MDS). The results of the rarefaction analysis that was conducted in this study suggested that between 161 and 259 OTUs (97% identity level) could be found in each of the 4 libraries for which full length 16S rRNA clones were constructed. This is in agreement with the earlier estimates of 177 OTUs (Edwards et al., 2004b). When all 16S rDNA libraries were compared, 510 unique OTUs were observed. It appeared that the most populous organism (64% of the sequences in 59 OTUs) was present in all the animals, contrasting with 10% of the sequences in 273 OTUs that were present in only 1 animal (Brulc et al., 2009). The microbial diversity in these 4 samples was correlated with the bacterial EGTs in the liquid and solid fractions of the samples that were analyzed (Brulc et al., 2009). The bacterial EGTs in these samples accounted for up to 95% of total EGTs. The bulk of these belonged to Bacteriodetes, Firmicutes, and Proteobacteria. Ruminococcus was rare in these sample sets. Archaeal species EGTs accounted for 2.3% of the total EGTs with other estimates of archaeal populations in the rumen. Only a few Eukaryotes were identified (1.3%) and no fungal rRNA gene sequences were identified. Viral EGTs were rare (0.1%). The microbiome samples from all animals had the presence of gene sequences to metabolize plant cellwall carbohydrates components including cellulose and xylan. A number of similarities and differences between this dataset and the previously documented termite and chicken GIT studies were presented in this study. For example, a wide variety of GH catalytic modules were observed in the rumen samples, but the termite hindgut contains more modules that would be associated with the degradation of cellulose and xylan, and as the authors propose, this is likely because of differences in diets being consumed by the different species (Warnecke et al., 2007). The termite gut microbiome also had more genes involved in N fixation (Warnecke et al., 2007). Compared to the Soudan Mine study, where two communities that were proximately close had distinctly different metabolisms, Dinsdale et al. (2008) found characteristic metabolisms in metagenomes that were similar environmentally. Although all metagenomes were functionally diverse (number of metabolic processes represented), the evenness or relative abundance of a particular metabolism depended on the biome origin indicating that the frequency of a gene encoding a particular metabolic function reflects its relative importance in an environment. RUMEN METAGENOMICS Limited varied studies have been conducted on the rumen metagenome, the most detailed of which today are those that have been conducted in our laboratory. In the earliest of these so-called metagenomic studies of ruminants, Ferrer et al. (2005) described “Novel hydrolase diversity retrieved from a metagenome library of bovine rumen microflora”. For this study, the rumen samples were derived from one New Zealand dairy cow and the metagenomic expression library created in a bacteriophage lambda ZAP phagemid vector (Stratagene). Inserts were screened for esterase, cellulose, and amylase-like activities. The 22 clones with hydrolytic capabilities (Ferrer et al., 2005) included 12 esterases, 9 endo-beta-1,4-glucanases, and 1 cyclodextrinase. Eight of these were being observed for the first time, and highlight the potential of the rumen for holding novel enzymatic and biochemical activities. In a subsequent study, this same rumen metagenomic library was used to identify RL5, a gene coding for a novel polyphenol oxidase whose recombinant protein revealed an ability to work on a range of substrates including 2,6dimethoxyphenol, tetramethylbenzidine, and phenol red (Beloqui et al., 2006). The signficance of this “laccase” was the lack of similarity to other previously described laccases and wide substrate affinity. Most recently, this group described a novel amylase from the same rumen metagenomic library. The enzyme had activity towards maltooligosaccharides, cyclomaltoheptaose, cyclomaltohexaose, cyclomaltooctaose, soluble starch, amylose, pullulan and amylopectin, and it was further characterized as being a cyclomaltodextrinase (Ferrer et al., 2007). Liu et al., (2009) described the isolation and characterization of lipases from a rumen metagenomic library. The rumen library was constructed from Holstein cows. One of the lipases showed highest similarity to a carboxylesterase from Thermosinus carboxydivorans Nor1. Both recombinant lipases were thermally unstable and had high activity to laurate, palmitate, and stearate (Liu et al., 2009). Studies with rumen samples from buffalo have also revealed novel genes that encode acidic cellulases (Duan et al., 2009). Here, the authors created metagenomic libraries and screened for functionalities to identify 61 independent clones with cellulase activity. We published the largest metagenomic gene set from ruminant species to date (Brulc et al., 2009). For this study, a combination of sequencing, assembly, phylotype, and subsystems-based analysis was used to study 15 in the rumen. The conversion of plant biomass to products of nutritional value is carried out by a consortium of microbes. The complex chemical processes involved in this process are not carried out by any one species. The differences in microbial abundance, and therefore metabolic contributions to the digestive tract, may indicate that highly represented bacteria are more significant contributors to nutrient utilization, animal health and wellbeing, and food safety; however, less abundant organisms are no less important because they may make significant and (or) unique metabolic contributions in these complex environments. A better understanding of the metabolic potential in the rumen microbial community and how we might utilize this potential is arising from metagenomics and offers possibilities for the discovery of novel metabolic pathways and enzymatic functions that could have profound implications for animal production. Moreover, rumen metagenomics will provide strategies that have significant environmental benefits, such as reduced fecal waste and a decrease in ruminant-derived methane production. Metagenomic-mediated information has the profound potential to increase feed efficiency, thereby reducing the amount of feed per animal and increasing the resulting products of commercial interest. As we continue to use metagenomic approaches to investigate the rumen, we will learn more about the previously uncultured species that reside there, their contributions to the microbiome and their potential to improve the overall ruminant farming process. By doing so, metagenomic approaches will contribute productively to animal growth and nutrient utilization and foster a positive image for the agricultural sector with respect to the environment. SUBTRACTIVE HYBRIDIZATION Another approach to identify functional genomic differences between rumen microbial communities, regardless of phylogenetic origin, is the use of suppressive subtractive hybridization (SSH, also referred to as suppression subtractive hybridization). Suppressive subtractive hybridization enables the identification of unique microbial DNA sequences present in one metagenomic sample relative to another, by subtracting the ones that are common to both, and amplifying the unique sequences. The technique was originally developed to amplify rare mRNA transcripts for full representation in mRNA libraries from eukaryotic tissues (Bonaldo et al., 1996). Suppressive subtractive hybridization introduced the use of specially designed adaptors. When these adaptors are ligated to themselves or to small DNA fragments then the secondary structure (a “panhandle”-like structure) introduced by their sequence content effectively inhibits the PCR amplification in the final preparation. This efficiently decreases the detection of background fragments and allows the identification of larger, differential DNA of sufficient “informational” length. The amended version of SSH was demonstrated to compare the genomes of two strains of Helicobacter pylori (Akopyants et al., 1998). We applied the SSH technique to a mixed microbial community to compare the metagenomes of microbiome from rumen samples (Galbraith et al., 2004). Our SSH study was conducted with samples of total genomic DNA obtained from rumen fluid samples of 2 steers fed hay. Tester-specific SSH fragments were found in 95 of 96 randomly selected clones. This metagenomic approach was able to expose an unexpectedly large difference in Archaeal community structure between the rumen microbial populations of 2 steers fed identical diets and housed together. This work was conducted using very small Sanger DNA sequence sets, and we anticipate that when applied to deep 454 pyrosequencing samples, that unique metabolisms and taxonomies will be revealed. LITERATURE CITED Akopyants, N. S., S. W. Clifton, D. Kersulyte, J. E. Crabtree, B. E. Youree, C. A. Reece, N. O. Bukanov, E. S. Drazek, B. A. Roe, and D. E. Berg. 1998. Analyses of the cag pathogenicity island of Helicobacter pylori. Mol. Microbiol. 28:37-53. Amann, R. I., W. Ludwig, and K. H. Schleifer. 1995. Phylogenetic identification and in situ detection of individual microbial cells without cultivation. Microbiol. Rev. 59:143-169. Attwood, G. T., K. Reilly, and B. K. Patel. 1996. Clostridium proteoclasticum sp. nov., a novel proteolytic bacterium from the bovine rumen. Int. J. Syst. Bacteriol. 46:753-758. Aurilia, V., J. C. Martin, S. I. McCrae, K. P. Scott, M. T. Rincon, and H. J. Flint. 2000. Three multidomain esterases from the cellulolytic rumen anaerobe Ruminococcus flavefaciens 17 that carry divergent dockerin sequences. Microbiol. 146:1391-1397. Baar, C., M. Eppinger, G. Raddatz, J. Simon, C. Lanz, O. Klimmek, R. Nandakumar, R. Gross, A. Rosinus, H. Keller, P. Jagtap, B. Linke, F. Meyer, H. Lederer, and S. C. Schuster. 2003. Complete genome sequence and analysis of Wolinella succinogenes. Proc. Natl. Acad. Sci. (USA) 100:11690-11695. SUMMARY The advent of next generation cost-effective highthroughput sequencing is allowing the application of metagenomics to the rumen microbiota. Metagenomics will allow us to combine an assessment of microbial population structure with an assessment of microbial community functionality. Although initial metagenomic studies demonstrated that there are differences in the microbiome of animals fed different diets, we have only begun to address other fundamental questions such as what microbial factors are associated with increased feed efficiency, nutrient utilization, improved animal health and well-being, and food safety. Animal feed is one of the largest costs in beef cattle production, and ruminants have the unique ability to derive nutrients from forages and lowquality roughages. This ability has nothing to do with the animal’s digestive enzymes, which inherently lack the ability to degrade the major components of plant biomass, and are totally dependent upon the symbiotic relationship between the host animal and microbial populations residing 16 Duan, C. J., L. Xian, G. C. Zhao, Y. Feng, H. Pang, X. L. Bai, J. L. Tang, Q. S. Ma, and J. X. Feng. 2009. Isolation and partial characterization of novel genes encoding acidic cellulases from metagenomes of buffalo rumens. J. Appl. Microbiol. 107:245-256. Edwards, J. E., N. R. McEwan, A. J. Travis, and R. J. Wallace. 2004. 16S rDNA library-based analysis of ruminal bacterial diversity. Antonie van Leeuwenhoek 86:263-281. Edwards, R. A., B. Rodriguez-Brito, L. Wegley, M. Haynes, M. Breitbart, D. M. Peterson, M. O. Saar, S. Alexander, E. C. Alexander, and F. Rohwer. 2006. Using pyrosequencing to shed light on deep mine microbial ecology. BMC Genomics 7:57-69. Edwards, R. A. and F. Rohwer. 2005. Viral metagenomics. Nat. Rev. Microbiol. 3:504-510. Ferrer, M., A. Beloqui, O. V. Golyshina, F. J. Plou, A. Neef, T. N. Chernikova, L. Fernandez-Arrojo, I. Ghazi, A. Ballesteros, K. Elborough, K. N. Timmis, and P. N. Golyshin. 2007. Biochemical and structural features of a novel cyclodextrinase from cow rumen metagenome. Biotechnol. J. 2:207-213. Ferrer, M., O. V. Golyshina, T. N. Chernikova, A. N. Khachane, D. Reyes-Duarte, V. A. Santos, C. Strompl, K. Elborough, G. Jarvis, A. Neef, M. M. Yakimov, K. N. Timmis, and P. N. Golyshin. 2005. Novel hydrolase diversity retrieved from a metagenome library of bovine rumen microflora. Environ. Microbiol. 7:1996-2010. Fillingham, I. J., P. A. Kroon, G. Williamson, H. J. Gilbert, and G. P. Hazlewood. 1999. A modular cinnamoyl ester hydrolase from the anaerobic fungus Piromyces equi acts synergistically with xylanase and is part of a multiprotein cellulose-binding cellulase-hemicellulase complex. Biochem. J. 343:215-224. Flint, H. J., J. Martin, C. A. McPherson, A. S. Daniel, and J. X. Zhang. 1993. A bifunctional enzyme, with separate xylanase and ß(1,3-1,4)-glucanase domains, encoded by the xynD gene of Ruminococcus flavefaciens. J. Bacteriol. 175:2943-2951. Fox, G. E., L. J. Magrum, W. E. Balch, R. S. Wolfe, and C. R. Woese. 1977. Classification of methanogenic bacteria by 16S ribosomal RNA characterization. Proc. Natl. Acad. Sci. (USA) 74:4537-4541. Galbraith, E. A., D. A. Antonopoulos, and B. A. White. 2004. Suppressive subtractive hybridization as a tool for identifying genetic diversity in an environmental metagenome: the rumen as a model. Environ. Microbiol. 6:928-937. Garcia-Vallve, S., A. Romeu, and J. Palau. 2000. Horizontal gene transfer of glycosyl hydrolases of the rumen fungi. Mol. Biol. Evol. 17:352-361. Gill, S. R., M. Pop, R. T. Deboy, P. B. Eckburg, P. J. Turnbaugh, B. S. Samuel, J. I. Gordon, D. A. Relman, C. M. Fraser-Liggett, and K. E. Nelson. 2006. Metagenomic analysis of the human distal gut microbiome. Science 312:1355-1359. Beloqui, A., M. Pita, J. Polaina, A. Martinez-Arias, O. V. Golyshina, M. Zumarraga, M. M. Yakimov, H. GarciaArellano, M. Alcalde, V. M. Fernandez, K. Elborough, J. M. Andreu, A. Ballesteros, F. J. Plou, K. N. Timmis, M. Ferrer, and P. N. Golyshin. 2006. Novel polyphenol oxidase mined from a metagenome expression library of bovine rumen: biochemical properties, structural analysis, and phylogenetic relationships. J. Biol. Chem. 281:22933-22942. Berg Miller, M. E., D. A. Antonopoulos, M. T. Rincon, M. Band, A. Bari, T. Akraiko, A. Hernandez, J. Thimmapuram, B. Henrissat, P. M. Coutinho, I. Borovok, S. Jindou, R. Lamed, H. J. Flint, E. A. Bayer, and B. A. White. 2009. Diversity and strain specificity of plant cell wall degrading enzymes revealed by the draft genome of Ruminococcus flavefaciens FD-1. PLoS One 4:e6650. Bonaldo, M. D. F., G. Lennon, and M. B. Soares. 1996. Normalization and subtraction: Two approaches to facilitate gene discovery. Genome Res. 6:791-806. Borneman, W. S., L. G. Ljungdahl, R. D. Hartley, and D. E. Akin. 1992. Purification and partial characterization of two feruloyl esterases from the anaerobic fungus Neocallimastix strain MC-2. Appl. Environ. Microbiol. 58:3762-3766. Bowman, B. H., J. W. Taylor, A. G. Brownlee, J. Lee, S. D. Lu, and T. J. White. 1992. Molecular evolution of the fungi: relationship of the Basidiomycetes, Ascomycetes, and Chytridiomycetes. Mol. Biol. Evol. 9:285-296. Brookman, J. L., G. Mennim, A. P. Trinci, M. K. Theodorou, and D. S. Tuckwell. 2000. Identification and characterization of anaerobic gut fungi using molecular methodologies based on ribosomal ITS1 and 185 rRNA. Microbiol. 146:393-403. Brulc, J. M., D. A. Antonopoulos, M. E. Miller, M. K. Wilson, A. C. Yannarell, E. A. Dinsdale, R. E. Edwards, E. D. Frank, J. B. Emerson, P. Wacklin, P. M. Coutinho, B. Henrissat, K. E. Nelson, and B. A. White. 2009. Gene-centric metagenomics of the fiber-adherent bovine rumen microbiome reveals forage specific glycoside hydrolases. Proc. Natl. Acad. Sci. (USA) 106:1948-1953. Demeyer, D. 1981. Rumen microbes and digestion of plant cell walls. Agr. Environ. 6:295 Denman, S. E. and C. S. McSweeney. 2006. Development of a real-time PCR assay for monitoring anaerobic fungal and cellulolytic bacterial populations within the rumen. FEMS Microbiol. Ecol. 58:572-582. Dethlefsen, L., S. Huse, M. L. Sogin, and D. A. Relman. 2008. The pervasive effects of an antibiotic on the human gut microbiota, as revealed by deep 16S rRNA sequencing. Plos Biol. 6:2383-2400. Dinsdale, E. A., R. A. Edwards, D. Hall, F. Angly, M. Breitbart, J. M. Brulc, M. Furlan, C. Desnues, M. Haynes, L. Li, L. McDaniel, M. A. Moran, K. E. Nelson, C. Nilsson, R. Olson, J. Paul, B. R. Brito, Y. Ruan, B. K. Swan, R. Stevens, D. L. Valentine, R. V. Thurber, L. Wegley, B. A. White, and F. Rohwer. 2008. Functional metagenomic profiling of nine biomes. Nature 452:629-632. 17 Koike, S., S. Yoshitani, Y. Kobayashi, and K. Tanaka. 2003. Phylogenetic analysis of fiber-associated rumen bacterial community and PCR detection of uncultured bacteria. FEMS Microbiol. Lett. 229:23-30. Krause, D. O., B. P. Dalrymple, W. J. Smith, R. I. Mackie, and C. S. McSweeney. 1999. 16S rDNA sequencing of Ruminococcus albus and Ruminococcus flavefaciens: design of a signature probe and its application in adult sheep. Microbiol. 145:1797-1807. Kurokawa, K., T. Itoh, T. Kuwahara, K. Oshima, H. Toh, A. Toyoda, H. Takami, H. Morita, V. K. Sharma, T. P. Srivastava, T. D. Taylor, H. Noguchi, H. Mori, Y. Ogura, D. S. Ehrlich, K. Itoh, T. Takagi, Y. Sakaki, T. Hayashi, and M. Hattori. 2007. Comparative metagenomics revealed commonly enriched gene sets in human gut microbiomes. DNA Res. 14:169-181. Kysela, D. T., C. Palacios, and M. L. Sogin. 2005. Serial analysis of V6 ribosomal sequence tags (SARST-V6): a method for efficient, high-throughput analysis of microbial community composition. Environ. Microbiol. 7:356-364. Lee, P. C., S. Y. Lee, S. H. Hong, and H. N. Chang. 2002. Isolation and characterization of a new succinic acidproducing bacterium, Mannheimia succiniciproducens MBEL55E, from bovine rumen. Appl. Microbiol. Biotechnol. 58:663-668. Ley, R. E., M. Hamady, C. Lozupone, P. J. Turnbaugh, R. R. Ramey, J. S. Bircher, M. L. Schlegel, T. A. Tucker, M. D. Schrenzel, R. Knight, and J. I. Gordon. 2008a. Evolution of mammals and their gut microbes. Science 320:1647-1651. Ley, R. E., C. A. Lozupone, M. Hamady, R. Knight, and J. I. Gordon. 2008b. Worlds within worlds: evolution of the vertebrate gut microbiota. Nature Rev. Microbiol. 6:776-788. Ley, R. E., D. A. Peterson, and J. I. Gordon. 2006. Ecological and evolutionary forces shaping microbial diversity in the human intestine. Cell 124:837-848. Lin, C., B. Flesher, W. C. Capman, R. I. Amann, and D. Stahl. 1994. Taxon specific hybridization probes for fiber-digesting bacteria suggest novel gut-associated Fibrobacter. Syst. Appl. Microbiol. 17:418-424. Lin, C., L. Raskin, and D. Stahl. 1997. Microbial community structure in gastrointestinal tracts of domestic animals: comparative analyses using rRNAtargeted oligonucleotide probes. FEMS Microbiol. Ecol. 22:281-294. Lin, C. and D. A. Stahl. 1995. Taxon-specific probes for the cellulolytic genus Fibrobacter reveal abundant and novel equine-associated populations. Appl. Environ. Microbiol. 61:1348-1351. Liu, K., J. Wang, D. Bu, S. Zhao, C. McSweeney, P. Yu, and D. Li. 2009. Isolation and biochemical characterization of two lipases from a metagenomic library of China Holstein cow rumen. Biochem. Biophys. Res. Commun. 385:605-611. Goldberg, S. M., J. Johnson, D. Busam, T. Feldblyum, S. Ferriera, R. Friedman, A. Halpern, H. Khouri, S. A. Kravitz, F. M. Lauro, K. Li, Y. H. Rogers, R. Strausberg, G. Sutton, L. Tallon, T. Thomas, E. Venter, M. Frazier, and J. C. Venter. 2006. A Sanger/pyrosequencing hybrid approach for the generation of high-quality draft assemblies of marine microbial genomes. Proc. Natl. Acad. Sci. (USA) 103:11240-11245. Guettler, M. V., D. Rumler, and M. K. Jain. 1999. Actinobacillus succinogenes sp. nov., a novel succinicacid-producing strain from the bovine rumen. Int. J. Syst. Bacteriol. 49 Pt. 1:207-216. Handelsman, J. 2004. Metagenomics: Application of genomics to uncultured microorganisms. Microbiol. Mol. Biol. Rev. 68:669-685. Hespell, R. B., D. E. Akin, and B. A. Dehority. 1997. Bacteria, fungi and protozoa of the rumen. Pages 59186 in Gastrointestinal Microbiology. Vol. 2. R. I. Mackie, B. A. White, and R. Isaacson, ed. Chapman and Hall, New York. Hong, S. H., J. S. Kim, S. Y. Lee, Y. H. In, S. S. Choi, J. K. Rih, C. H. Kim, H. Jeong, C. G. Hur, and J. J. Kim. 2004. The genome sequence of the capnophilic rumen bacterium Mannheimia succiniciproducens. Nat Biotechnol. 22:1275-1281. Hugenholtz, P., B. M. Goebel, and N. R. Pace. 1998. Impact of culture-independent studies on the emerging phylogenetic view of bacterial diversity. J. Bacteriol. 180:4765-4774. Hugenholtz, P. and N. R. Pace. 1996. Identifying microbial diversity in the natural environment: a molecular phylogenetic approach. Trends Biotechnol. 14:190-197. Huse, S. M., L. Dethlefsen, J. A. Huber, D. M. Welch, D. A. Relman, and M. L. Sogin. 2008. Exploring Microbial Diversity and Taxonomy Using SSU rRNA Hypervariable Tag Sequencing. Plos Genetics 4:e1000255. Hyman, E. D. 1988. A New Method of Sequencing DNA. Anal. Biochem. 174:423-436. Janssen, P. H. and M. Kirs. 2008. Structure of the archaeal community of the rumen. Appl. Environ. Microbiol. 74:3619-3625. Jindou, S., I. Borovok, M. T. Rincon, H. J. Flint, D. A. Antonopoulos, M. E. Berg, B. A. White, E. A. Bayer, and R. Lamed. 2006. Conservation and divergence in cellulosome architecture between two strains of Ruminococcus flavefaciens. J. Bacteriol. 188:79717976. Kocherginskaya, S. A., R. I. Aminov, and B. A. White. 2001. Analysis of the rumen bacterial diversity under two different diet conditiions using denaturing gradient gel electrophoresis, random sequencing, and statistical ecology approaches. Anaerobe 7:119-134. Koike, S. and Y. Kobayashi. 2001. Development and use of competitive PCR assays for the rumen cellulolytic bacteria: Fibrobacter succinogenes, Ruminococcus albus and Ruminococcus flavefaciens. FEMS Microbiol. Lett. 204:361-366. 18 Qi, M., K. E. Nelson, S. C. Daugherty, W. C. Nelson, I. R. Hance, M. Morrison, and C. W. Forsberg. 2008. Genomic differences between Fibrobacter succinogenes S85 and Fibrobacter intestinalis DR7, identified by suppression subtractive hybridization. Appl. Environ. Microbiol. 74:987-993. Qin, J., R. Li, J. Raes, M. Arumugam, K. S. Burgdorf, C. Manichanh, T. Nielsen, N. Pons, F. Levenez, T. Yamada, D. R. Mende, J. Li, J. Xu, S. Li, D. Li, J. Cao, B. Wang, H. Liang, H. Zheng, Y. Xie, J. Tap, P. Lepage, M. Bertalan, J. M. Batto, T. Hansen, D. Le Paslier, A. Linneberg, H. B. Nielsen, E. Pelletier, P. Renault, T. Sicheritz-Ponten, K. Turner, H. Zhu, C. Yu, M. Jian, Y. Zhou, Y. Li, X. Zhang, N. Qin, H. Yang, J. Wang, S. Brunak, J. Dore, F. Guarner, K. Kristiansen, O. Pedersen, J. Parkhill, J. Weissenbach, M. Antolin, F. Artiguenave, H. Blottiere, N. Borruel, T. Bruls, F. Casellas, C. Chervaux, A. Cultrone, C. Delorme, G. Denariaz, R. Dervyn, M. Forte, C. Friss, M. van de Guchte, E. Guedon, F. Haimet, A. Jamet, C. Juste, G. Kaci, M. Kleerebezem, J. Knol, M. Kristensen, S. Layec, K. Le Roux, M. Leclerc, E. Maguin, R. Melo Minardi, R. Oozeer, M. Rescigno, N. Sanchez, S. Tims, T. Torrejon, E. Varela, W. de Vos, Y. Winogradsky, E. Zoetendal, P. Bork, and S. D. Ehrlich. 2010. A human gut microbial gene catalogue established by metagenomic sequencing. Nature 464:59-65. Rincon, M. T., T. Cepeljnik, J. C. Martin, R. Lamed, Y. Barak, E. A. Bayer, and H. J. Flint. 2005. Unconventional mode of attachment of the Ruminococcus flavefaciens cellulosome to the cell surface. J. Bacteriol. 187:7569-7578. Rodriguez-Brito, B., F. Rohwer, and R. A. Edwards. 2006. An application of statistics to comparative metagenomics. BMC Bioinformatics 7:162-173. Roesch, L. F. W., R. R. Fulthorpe, A. Riva, G. Casella, A. K. M. Hadwin, A. D. Kent, S. H. Daroub, F. A. O. Camargo, W. G. Farmerie, and E. W. Triplett. 2007. Pyrosequencing enumerates and contrasts soil microbial diversity. ISME J. 1:283-290. Ronaghi, M., S. Karamohamed, B. Pettersson, M. Uhlen, and P. Nyren. 1996. Real-time DNA sequencing using detection of pyrophosphate release. Anal. Biochem. 242:84-89. Ronaghi, M., M. Uhlen, and P. Nyren. 1998. A sequencing method based on real-time pyrophosphate. Science 281:363-365. Rondon, M. R., P. R. August, A. D. Bettermann, S. F. Brady, T. H. Grossman, M. R. Liles, K. A. Loiacono, B. A. Lynch, I. A. MacNeil, C. Minor, C. L. Tiong, M. Gilman, M. S. Osburne, J. Clardy, J. Handelsman, and R. M. Goodman. 2000. Cloning the soil metagenome: a strategy for accessing the genetic and functional diversity of uncultured microorganisms. Appl. Environ. Microbiol. 66:2541-2547. Schloss, P. D. and J. Handelsman. 2007. The last word: books as a statistical metaphor for microbial communities. Annu. Rev. Microbiol. 61:23-34. Margulies, M., M. Egholm, W. E. Altman, S. Attiya, J. S. Bader, L. A. Bemben, J. Berka, M. S. Braverman, Y. J. Chen, Z. Chen, S. B. Dewell, L. Du, J. M. Fierro, X. V. Gomes, B. C. Godwin, W. He, S. Helgesen, C. H. Ho, G. P. Irzyk, S. C. Jando, M. L. Alenquer, T. P. Jarvie, K. B. Jirage, J. B. Kim, J. R. Knight, J. R. Lanza, J. H. Leamon, S. M. Lefkowitz, M. Lei, J. Li, K. L. Lohman, H. Lu, V. B. Makhijani, K. E. McDade, M. P. McKenna, E. W. Myers, E. Nickerson, J. R. Nobile, R. Plant, B. P. Puc, M. T. Ronan, G. T. Roth, G. J. Sarkis, J. F. Simons, J. W. Simpson, M. Srinivasan, K. R. Tartaro, A. Tomasz, K. A. Vogt, G. A. Volkmer, S. H. Wang, Y. Wang, M. P. Weiner, P. Yu, R. F. Begley, and J. M. Rothberg. 2005. Genome sequencing in microfabricated high-density picolitre reactors. Nature 437:376-380. Moon, C. D., D. M. Pacheco, W. J. Kelly, S. C. Leahy, D. Li, J. Kopecny, and G. T. Attwood. 2008. Reclassification of Clostridium proteoclasticum as Butyrivibrio proteoclasticus comb. nov., a butyrateproducing ruminal bacterium. Int. J. Syst. Evol. Microbiol. 58:2041-2045. Morrison, M., S. C. Daugherty, W. C. Nelson, T. Davidsen, and K. E. Nelson. 2009. The FibRumBa Database: A Resource for Biologists with Interests in Gastrointestinal Microbial Ecology, Plant Biomass Degradation, and Anaerobic Microbiology. Microb. Ecol. 59:212-213. Neufeld, J. D. and W. W. Mohn. 2005. Unexpectedly high bacterial diversity in arctic tundra relative to boreal forest soils, revealed by serial analysis of ribosomal sequence tags. Appl. Environ. Microbiol. 71:57105718. Neufeld, J. D., Z. T. Yu, W. Lam, and W. W. Mohn. 2004. Serial analysis of ribosomal sequence tags (SARST): a high-throughput method for profiling complex microbial communities. Environ. Microbiol. 6:131-144. Nicholson, M. J., M. K. Theodorou, and J. L. Brookman. 2005. Molecular analysis of the anaerobic rumen fungus Orpinomyces - insights into an AT-rich genome. Microbiol. 151:121-133. Overbeek, R., T. Begley, R. M. Butler, J. V. Choudhuri, H. Y. Chuang, M. Cohoon, V. de Crecy-Lagard, N. Diaz, T. Disz, R. Edwards, M. Fonstein, E. D. Frank, S. Gerdes, E. M. Glass, A. Goesmann, A. Hanson, D. Iwata-Reuyl, R. Jensen, N. Jamshidi, L. Krause, M. Kubal, N. Larsen, B. Linke, A. C. McHardy, F. Meyer, H. Neuweger, G. Olsen, R. Olson, A. Osterman, V. Portnoy, G. D. Pusch, D. A. Rodionov, C. Ruckert, J. Steiner, R. Stevens, I. Thiele, O. Vassieva, Y. Ye, O. Zagnitko, and V. Vonstein. 2005. The subsystems approach to genome annotation and its use in the project to annotate 1000 genomes. Nucleic Acids Res. 33:5691-5702. Pace, N. R. 1997. A molecular view of microbial diversity and the biosphere. Science 276:734-740. 19 Veira, D. M. 1986. The role of ciliate protozoa in nutrition of the ruminant. J. Anim. Sci. 63:1547-1560. Venter, J. C., K. Remington, J. F. Heidelberg, A. L. Halpern, D. Rusch, J. A. Eisen, D. Wu, I. Paulsen, K. E. Nelson, W. Nelson, D. E. Fouts, S. Levy, A. H. Knap, M. W. Lomas, K. Nealson, O. White, J. Peterson, J. Hoffman, R. Parsons, H. Baden-Tillson, C. Pfannkoch, Y. H. Rogers, and H. O. Smith. 2004. Environmental genome shotgun sequencing of the Sargasso Sea. Science 304:66-74. Wallace, R. J., L. C. Chaudhary, N. McKain, N. R. McEwan, A. J. Richardson, P. E. Vercoe, N. D. Walker, and D. Paillard. 2006. Clostridium proteoclasticum: a ruminal bacterium that forms steric acid from linoleic acid. FEMS Microbiol. Lett. 265: 195–201. Warnecke, F., P. Luginbuhl, N. Ivanova, M. Ghassemian, T. H. Richardson, J. T. Stege, M. Cayouette, A. C. McHardy, G. Djordjevic, N. Aboushadi, R. Sorek, S. G. Tringe, M. Podar, H. G. Martin, V. Kunin, D. Dalevi, J. Madejska, E. Kirton, D. Platt, E. Szeto, A. Salamov, K. Barry, N. Mikhailova, N. C. Kyrpides, E. G. Matson, E. A. Ottesen, X. Zhang, M. Hernandez, C. Murillo, L. G. Acosta, I. Rigoutsos, G. Tamayo, B. D. Green, C. Chang, E. M. Rubin, E. J. Mathur, D. E. Robertson, P. Hugenholtz, and J. R. Leadbetter. 2007. Metagenomic and functional analysis of hindgut microbiota of a wood-feeding higher termite. Nature 450:560-565. Whitford, M. F., R. J. Forster, C. E. Beard, J. Gong, and R. M. Teather. 1998. Phylogenetic analysis of rumen bacteria by comparative sequence analysis of cloned 16S rRNA genes. Anaerobe 4:153-163. Woese, C. R. and G. E. Fox. 1977. Phylogenetic structure of the prokaryotic domain: the primary kingdoms. Proc. Natl. Acad. Sci. (USA) 74:5088-5090. Woese, C. R., O. Kandler, and M. L. Wheelis. 1990. Towards a natural system of organisms: proposal for the domains Archaea, Bacteria, and Eucarya. Proc. Natl. Acad. Sci. (USA) 87:4576-4579. Yu, Z., M. Yu, and M. Morrison. 2006. Improved serial analysis of V1 ribosomal sequence tags (SARST-V1) provides a rapid, comprehensive, sequence-based characterization of bacterial diversity and community composition. Environ. Microbiol. 8:603-611. Zhang, J. X. and H. J. Flint. 1992. A bifunctional xylanase encoded by the xynA gene of the rumen cellulolytic bacterium Ruminococcus flavefaciens 17 comprises two dissimilar domains linked by an asparagine/glutaminerich sequence. Mol. Microbiol. 6:1013-1023. Ziemer, C. J., R. Sharp, M. D. Stern, M. A. Cotta, T. R. Whitehead, and D. A. Stahl. 2000. Comparison of microbial populations in model and natural rumens using 16S ribosomal RNA-targeted probes. Environ. Microbiol. 2:632-643. Schloss, P. D. and J. Handelsman. 2008. A statistical toolbox for metagenomics: assessing functional diversity in microbial communities. BMC Bioinformatics 9:34. Shinkai, T. and Y. Kobayashi. 2007. Localization of ruminal cellulolytic bacteria on plant fibrous materials as determined by fluorescence in situ hybridization and real-time PCR. Appl. Environ. Microbiol. 73:16461652. Sogin, M. L., H. G. Morrison, J. A. Huber, D. M. Welch, S. M. Huse, P. R. Neal, J. M. Arrieta, and G. J. Herndl. 2006. Microbial diversity in the deep sea and the underexplored "rare biosphere". Proc. Natl. Acad. Sci. (USA) 103:12115-12120. Stahl, D. A., B. Flesher, H. R. Mansfield, and L. Montgomery. 1988. Use of phylogenetically based hybridization probes for studies of ruminal microbial ecology. Appl. Environ. Microbiol. 54:1079-1084. Tajima, K., R. I. Aminov, T. Nagamine, H. Matsui, M. Nakamura, and Y. Benno. 2001a. Diet-dependent shifts in the bacterial population of the rumen revealed with real-time PCR. Appl. Environ. Microbiol. 67:27662774. Tajima, K., R. I. Aminov, T. Nagamine, K. Ogata, M. Nakamura, H. Matsui, and Y. Benno. 1999. Rumen bacterial diversity as determined by sequence analysis of 16S rDNA libraries. FEMS Microbiol. Ecol. 29:159169. Tajima, K., S. Arai, K. Ogata, T. Nagamine, H. Matsui, M. Nakamura, R. I. Aminov, and Y. Benno. 2000. Rumen bacterial community transition during adaptation to high-grain diet. Anaerobe 6:273-284. Tajima, K., T. Nagamine, H. Matsui, M. Nakamura, and R. Aminov. 2001b. Phylogenetic analysis of archaeal 16S rRNA libraries from the rumen suggests the existence of a novel group of archaea not associated with known methanogens. FEMS Microbiol. Ecol. 200:67-72. Tringe, S. G., C. von Mering, A. Kobayashi, A. A. Salamov, K. Chen, H. W. Chang, M. Podar, J. M. Short, E. J. Mathur, J. C. Detter, P. Bork, P. Hugenholtz, and E. M. Rubin. 2005. Comparative metagenomics of microbial communities. Science 308:554-557. Turnbaugh, P. J., M. Hamady, T. Yatsunenko, B. L. Cantarel, A. Duncan, R. E. Ley, M. L. Sogin, W. J. Jones, B. A. Roe, J. P. Affourtit, M. Egholm, B. Henrissat, A. C. Heath, R. Knight, and J. I. Gordon. 2009. A core gut microbiome in obese and lean twins. Nature 457:480-484. Turnbaugh, P. J., R. E. Ley, M. Hamady, C. M. FraserLiggett, R. Knight, and J. I. Gordon. 2007. The human microbiome project. Nature 449:804-810. Turnbaugh, P. J., R. E. Ley, M. A. Mahowald, V. Magrini, E. R. Mardis, and J. I. Gordon. 2006. An obesityassociated gut microbiome with increased capacity for energy harvest. Nature 444:1027-1031. Tyson, G. W., J. Chapman, P. Hugenholtz, E. E. Allen, R. J. Ram, P. M. Richardson, V. V. Solovyev, E. M. Rubin, D. S. Rokhsar, and J. F. Banfield. 2004. Community structure and metabolism through reconstruction of microbial genomes from the environment. Nature 428:37-43. 20 21 Incomplete Closed Bovine Bovine Bovine Elephant Bovine Bovine Bovine Bovine Bovine Bovine Mannheimia succiniproducens Methanobrevibacter ruminantium M1 Orpinomyces OUS1 Piromyces E2 Prevotella ruminicola 23 Prevotella bryantii B14 Ruminococcus albus 8 Ruminococcus flavefaciens FD-1 Selenomonas ruminantium subsp. lactilytica TAM642 Wolinella succinogenes DSM1740 Draft (>10x) Draft (8x) Closed In progress In progress Closed Closed Closed 2.11 n/a 4.4 4.2 3.8 3.62 n/a n/a 2.9 2.31 3.84 4.14 Genome size (Mb) 2.32 48.5 n/a 45 n/a 39 48 ~15-20 ~15-20 33 42.5 48 40 % GC content 46 Joint Genome Institute, California, USA 2 AgResearch Limited, Palmerston North, New Zealand 3 The North American Consortium for Genomics of Fibrolytic Rumen Bacteria 4 Korea Advanced Institute of Science and Technology, Daejeon, Republic of Korea 5 Pastoral Greenhouse Gas Research Consortium, Wellington, New Zealand 6 Pacific Northwest National Laboratory, Washington, USA 7 National Institute of Technology and Evaluation, Tokyo, Japan 8 MaxPlanck Institute, Tübingen, Germany 1 Draft (29x) Bovine Fibrobacter succinogenes S85 Closed Bovine Butyrivibrio proteoclasticus B316 Closed Status Bovine Origin Actinobacillus succinogenes 130Z Species Table 1. Rumen microbial genome sequencing initiatives 5 MPI BX571656 E n/a NITE7 8 ACOK00000000G NACGFRB3 FibRumBa F 3 NACGFRB FibRumBaF NACGFRB3 n/a FibRumBaF 1 n/a CP001719 F NACGFRB3 PNWNL / JGI 6 PNWNL6 / JGI1 AgResearch / PGGRC 2 AE016827G KAIST4 n/a FibRumBaF NACGFRB3 AgResearch CP000746G JGI1 2 Available Sequencing institute Baar et al. (2003) n/a Berg Miller et al. (2009) n/a Purushe et al. (unpublished data) Purushe et al. (unpublished data) n/a n/a Leahy et al. (2010) Hong et al. (2004) Morrison et al. (unpublished data) Kelly et al. (unpublished data) n/a Reference EMERGING METHODS IN RUMEN MICROBIOLOGY J. M. Brulc, C. J. Yeoman, K. E. Nelson, and B. A. White Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 PRACTICAL APPLICATION OF MODERN RUMEN MICROBIOLOGICAL TECHNIQUES TO GRAZING RUMINANTS S. L. Lodge-Ivey1 Department of Animal and Range Science, New Mexico State University, Las Cruces 88003 (4 yr average = approximately $205/cow; Miller et al., 2001). Therefore, increasing efficiency of forage utilization by grazing ruminants will reduce feed costs and augment sustainable rangeland agriculture production systems because ruminants are uniquely capable of converting rangeland forage into consumable products. Domestication of herbivores dates back as far as 10,000 years ago, and herbivores still provide products that are useful in our daily lives. Cattle, sheep, and goats have the ability to convert forages into high value protein products for human use and thereby convert lands unsuitable for crop production to usable lands. Over the past 50 years, ruminant animal production on grazing lands has become more efficient due to improvements in genetic selection, as well as increased knowledge of feed utilization and land management. Improvements in production of grazing ruminants resulting directly from manipulations of the rumen microbial ecology have yet to come to fruition. Supplementation of protein and energy has proven to be a useful tool in increasing performance by meeting nutrient deficiencies inherent in the forage base. The increased production efficiency due to supplementation is related to changes in VFA profile and ruminal NH3-N production, which likely result in changes of the microbial population dynamics. Although the application of modern tools such as pyrosequencing, metagenomics, and the many possible combinations of next-generation sequencing technologies have allowed for an increased understanding of microbial diversity and detection of novel genes involved in nutrient processing (Brulc et al., 2010), direct linkages between shifts in microbial ecology that can be detected with sophisticated molecular biology techniques and animal performance are limited. The objective of this paper is to share ideas germane to the application of modern techniques used to study rumen microbial ecology of grazing ruminants, including techniques for sampling and selection of appropriate modern methodology for a given research question. ABSTRACT: Domestication of herbivore species dates back as far as 10,000 years ago, and herbivores today still provide products that are useful in our daily lives. Production of ruminants has changed over the past 50 years due to improvements in genetic selection and technical advancements in feeding and land management practices. Most ruminant animal production systems around the world use fibrous feeds as the bulk portion of ruminant diets. High levels of forage in ruminant diets results in less efficient production compared with feeding diets with greater levels of concentrate. Further characterization of the rumen microbial communities may eventually lead to enhanced production efficiency of grazing ruminants. Conventional culture-based methods of enumerating rumen microorganism are being replaced by culture independent methods that rely on analysis of nucleic acids extracted from ruminal samples. Sampling location within the rumen, liquid vs. particulate fraction, and sample handling are all areas of consideration when collecting ruminal samples for analysis with DNA- and RNA-based techniques. This paper is intended to provide suggestions for methods to collect rumen samples from grazing ruminants and selection of molecular techniques to answer questions about rumen microbial communities relevant to grazing ruminant production systems. INTRODUCTION The majority of ruminant animal production systems around the world incorporate fibrous feeds into diets. The US has approximately 1.8 billion acres of land deemed suitable for food and fiber production and 74% of this area is classified as forest, range, or pasture land (FAO, 2009). Approximately 50% of the land area in North America is rangeland suitable for grazing (Rinehart, 2008). Almost half of the global carbon fixed annually by photosynthesis is incorporated into plant cell walls making rangeland forage the most renewable carbon source on the planet (Krause et al., 2003). Rangelands and grasslands have the capability of producing millions of tons of energy annually (Rinehart, 2008). Rate of gain and feed efficiency are less when ruminant animals are fed high levels of forage vs. being fed high levels of cereal grains. Additionally, supplemental feed costs are the greatest factor influencing profit of commercial beef cows operations, accounting for over 63% of the variation in total annual cow costs RUMEN SAMPLING Location Obtaining quality data from the application modern techniques used to study rumen microbial ecology is dependent upon collecting a representative sample of ruminal contents. The composition of the rumen microflora is known to be responsive to changes in diet, antibiotic usage, and age and health of host animal, and varies with 1 Corresponding author: [email protected] 24 season, geographical location and feeding regimen (Kocherginskaya et al., 2001). Ruminal samples can be obtained via ruminal cannula, oral lavage, or at slaughter. Choice of ruminal sampling method and frequency of sampling grazing ruminants are complicated by the fact that experimental animals are not housed in pens and may be grazing in pastures encompassing very large areas. Potential impact that sample collection could have on animal foraging behavior, pasture utilization, loss of animal value, and limitation of experimental animals due to rumen cannula implantation should also be considered during the study design phase. Sample handling after collection is another consideration due to the potential remoteness of pasture locations and animal location within the pasture coupled with the need to obtain intact high quality DNA, and in some cases, RNA. Sampling location within the rumen has been observed to influence ruminal pH, VFA, and NH3-N (Leedle et al., 1982); however, few publications have compared sampling location using molecular biology techniques. Li et al. (2009) used ruminally cannulated dairy cows fed a lactation diet and collected ruminal fluid from 5 locations in the rumen at 3 time points. Ruminal fluid was strained and analyzed for VFA, pH, NH3-N, and microbial diversity using PCR-denaturing gradient electrophoresis (DGGE). Those authors reported that detectable bacterial structure was highly conserved among different locations in the rumen and, while the quantity of individual bacterial species may change over time, the changes were not correlated with pH, VFA, or NH3-N. The observations by Li et al. (2009) are in contrast to Leedle et al. (1982), who showed that total bacteria counts and fermentation acids increased after feeding. This represents a fundamental difference between culture studies and some DGGE studies. Leedle et al. (1982) used different carbohydrate medium and replicate plating techniques to assess diurnal variation in direct and viable cell counts in cattle fed high-forage or high-concentrate. Li et al. (2009) quantified total bacteria using DGGE, which relied on DNA extracted from ruminal fluid and a PCR primer set for amplification of the V2-V3 region of the16S rDNA gene. The similarity analysis relied on quantifying total number of bands in each sample and compared the resulting patterns to the other samples. The total number of bands is related to the number of dominant phenotypes (Muyzer et al., 1993) and this type of bacterial quantification method does not take into account the density of each bacterial species. Density of bacterial species can be quantified by analyzing DGGE data for band intensity or quantitative real-time PCR (qRT-PCR). Real-time PCR was used by Li et al. (2009) to determine bacterial density of 8 species of ruminal bacteria and confirmed that the quantity of individual bacterial species may change over time; however, this was not correlated with ruminal pH, VFA, or NH3-N. These data were consistent with those of Tajima et al. (2001), who pooled samples from a number of locations within the rumen of dry dairy cows and used qRT-PCR to quantify 13 bacterial species of rumen bacteria during abrupt dietary transitions from hay to a grain diet. Tajima et al. (2001) observed that cellulolytic bacteria decreased 10- fold during diet transition and amylolytic bacteria increased 7 to 375-fold during the same diet transition. Collection samples from ruminants via oral lavage to characterize rumen microbial ecology has been used in a number of studies. Oral lavage allows for sampling of a greater number of animals including wildlife and grazing ruminants with the added advantage that the value of the animals is not lost due to the surgical implantation of a ruminal cannula. The drawback is the inability to sample the rumen from a variety of locations. However, LodgeIvey et al. (2009) reported that collecting ruminal samples by oral lavage vs. rumen cannula did not affect DGGE profiles, VFA, or NH3-N. Collecting ruminal samples at slaughter is commonly used to study the gut contents of wildlife and feedlot animals. Samples collected at slaughter have been used to characterize the microbial diversity of the rumen microbiome of Norwegian reindeer (Sundset et al., 2009), two geographically separated sub-species of reindeer (Sundset et al., 2007), and feedlot steers (Guan et al., 2008). Ruminal contents should be collected immediately after slaughter (Sundset et al., 2007, 2009; Guan et al., 2008; Li et al., 2009), particularly if pH, VFA, and NH3-N are to be analyzed because these values could be inflated as a result of reduced transport of metabolites out of the rumen when the animal is slaughtered. Particulate vs. Liquid Fraction The question of which fraction of the rumen material should be sampled is related to sampling location, but is further refined to address sampling the rumen liquor, fiber particulates, or whole rumen fluid fractions. Results from electron microscopy indicate that there are a large number of bacteria, protozoa, and fungi attached to undigested feed particles in the rumen. Forsberg and Lam (1977) estimated that 75% of microbial ATP was associated with fiber particulate fraction. Likewise, Craig et al. (1987) observed that 70 to 80% of microbial organic matter in whole rumen contents was associated with the particulate phase and that particulate-associated microbial organic matter was greatest within 1 h after feeding. Animal Variation Animal variation may be a contributor to bacterial diversity noted when using molecular biology techniques. Microbial communities present in the rumen are very complex and individual animals fed similar diets and housed together may have differences in their rumen microbial community (Kocherginskaya et al., 2001; Regensbogenova et al., 2004). Regensbogenova et al. (2004) observed animal-to-animal variation when quantifying ruminal ciliates using DGGE, but no difference was observed when samples were collected from single sheep at differing time points. Individual animal seemed to impact DGGE grouping rather than species or sampling method (Lodge-Ivey et al., 2009). Likewise, Li et al. (2009) noted low similarity of detectable bacterial diversity among rumen digesta samples collected from different animals fed the same diet. However, diet seems to affect the magnitude 25 structure from genomic DNA-based analysis does not take into account the viability or metabolic state of the community members. In order to describe metabolic activity of a microbial community, RNA-based methods are more suitable (Kang et al., 2009). The production of rRNA produced by microbial cells is directly correlated with the growth activity of bacteria (Wagner, 1994). Ribonucleic acid is much more susceptible to attack by nuclease than DNA and enzymatic degradation of RNA is probably the greatest source of loss encountered with RNA isolation procedures. In addition to the precautions outlined for DNA isolation, samples for RNA extraction should also be collected into to containers treated with diethylpyrocarbonate (DEPC) and all reagents and glassware to be used during RNA isolation should be treated with a 0.1% DEPC solution followed by autoclaving. Also, all samples to be used for RNA isolation should be frozen and stored at -80o C. Kang et al. (2009) evaluated 4 methods to extract RNA for ruminal and fecal samples for efficiency and purity of RNA. It was found that treating samples with a dissociation solution followed by RNA extraction with Trizol after bead beating produced a higher quality and quantity of RNA when compared with similar methods using phenol/chloroform. of animal-to-animal variation detected among rumen microbial communities. Kocherginskaya et al. (2001) found that steers fed a 70% concentrate diet varied more from one another than cattle consuming 70% medium-quality alfalfa hay diet. The above observations also could be related the limited number of animals (typically 2 to 4 per treatment) sampled for rumen microbial community characterization. Thus, it is important to recognize that it is often difficult to separate animal variation from experimental variation due to the small number of animals typically used to characterize rumen microbial ecology. Sampling of few animals is common and is most likely related to cost of sample analysis (e.g., refer to Brulc et al., 2010). As molecular techniques for studying rumen microbial ecology evolve, sampling of just a handful of animals will not be adequate to answer questions related to the interaction of the rumen microbial population and animal performance. Fortunately, technologies such as pyrosequencing are now available which provide highthroughput capacity and continuous sequence reads from 250-500 bp in a cost effective manner (McSweeney et al., 2009). Greater cost savings is also achieved with phylogenetic based studies where researchers can take advantage of “barcoding” individual samples by incorporating unique base pair tags into the PCR primers for each sample then pooling the samples before sequencing. This allows for greater statistical power in experimental design and data interpretation where previously individual samples were pooled before sequencing. Kunin et al. (2008) explained that even though next generation sequencing allows for metagenomic analysis of complex microbiomes leading to phylogenetic descriptions of the system, it is still difficult to assign functional roles of individual bacteria within a metagenome due to small read lengths and the complexity of the community. APPLICATION OF TECHNIQUES Areas of interest for researchers studying grazing animals may include influence of season (diet quality), nutrient supplementation, impact of ingestion of toxic plants, surveying rumen microbial ecology of efficient vs. inefficient grazing ruminants, and determination of microbial ecology of wildlife. Nucleic acid-based technologies can now be employed to examine microbial community structure primarily through the use of small subunit (SSU) rRNA gene analysis including 16S and 18S rRNA and functioning of complex microbial ecosystems can be described by combined analysis of multiple genomes (metagenomics) and message from expressed gene (mRNA) in the rumen. Before choosing a molecular technique for studying the rumen microbial ecology of grazing ruminants, the researcher must decide if the goal is to quantify, indentify, or monitor microbial communities and (or) gene function because this will determine the techniques or combination of techniques that should be utilized. Also, common ruminant nutrition indicator analyses, such as ruminal pH, VFA, and NH3-N, and animal production responses should be included in microbial ecology experiments. Because most nucleic acid-based techniques rely on bulk-extracted nucleic acids, the assumption is that recovered 16S rRNA sequences relate to an in vivo active member of the microbial community. That may not always be the case and shifts in microbial communities may be interesting, and certainly within detection sensitivity of many current techniques, but until there is linkage between microbial communities and animal performance, some of these data will not contribute to enhancing production efficiency of grazing ruminants. An example of this was summarized by McSweeney et al. (2009) where Proteobacteria had commonly been believed to make up less than 2% of the Sampling Handling after Collection The steps following sample collection are crucial to obtaining high-quality DNA or RNA. Enzymes responsible for the degradation of RNA and DNA are ubiquitous in most environments as well as in microbial cells and may be responsible for the significant loss of nucleic acids during purification. Standard precautions such as wearing gloves while collecting samples, immediate chilling of ruminal samples by placing samples on ice, dry ice, or in liquid N and autoclaving of reagents will be adequate to preserve DNA for later extraction. Samples for DNA extraction can either be extracted immediately or frozen at -20oC for later processing. Isolation of nucleic acids from ruminal fluid is not a trivial matter; therefore, extraction of community DNA representing the complete diversity of rumen microbial communities is required. Methods that increase DNA yield will in theory translate to better representation of community DNA. Yu and Morrison (2004) reported a repeated bead-beating plus column extraction method that is commonly used by most laboratories today. Generally, DNA from rumen digesta can be used as a template for PCR amplification of 16S rRNA genes and subsequent community analysis. However, inferred community 26 uncultured bacteria, thereby making isolation more efficient. A number of fibrolytic bacteria in the rumen are uncultured. For example, 77% of the fiber-associated bacteria community members are uncultured (Koike et al., 2003). Koike et al. (2010) focused on two groups of uncultured fiber-associated bacterial groups, unculutured group 2 and 3 (U2 and U3, respectively). Those authors used qRT-PCR and fluorescent in situ hybridization (FISH) to characterize members of U2 and U3. To enrich for members of U2, rice straw or cellulose was incubated in the rumen in nylon bags. After 5 days, samples incubated in the rumen were inoculated into enriching medium described by Bryant and Burkey (1953). The enrichment was then serially diluted and transferred to agar containing rice straw. Isolated colonies were picked and characterized using PCR. Koike et al. (2010) were able to isolate two strains of bacteria that seem to be important to hemicellulose degradation in the rumen. This research is an example of how nucleic acid-based techniques can be combined with traditional culturing methods to attain a greater understanding of the rumen microbial community. These techniques may also be useful to isolate bacteria involved in toxic plant degradation in the rumen. Similar techniques used to study bacterial communities in the rumen can also be utilized to study protozoal and fungal populations. Techniques such as qRT-PCR and DGGE have been used to monitor the abundance and diversity of several genera and species of rumen protozoa (see McSweeney et al., 2006 for extensive review). Anaerobic fungi in the rumen also can be studied by qRTPCR (Edwards et al., 2008); however, qRT-PCR does not discriminate between the fungal genera and is therefore not useful to characterize diversity. Edwards et al. (2008) proposed that rumen fungal community characterization is better studied using the technique of automated ribosomal intergenic spacer region (ARISA). Furthermore, study of temporal interrelations between diets and different members of the rumen fungi population was possible with ARISA (Edwards et al., 2008). microbial community based on rDNA analysis. Analysis based on RNA showed that Proteobacteria represented 28% of the clones in a 16S rDNA library. Members of Proteobacteria group include a variety of pathogenic bacteria (Escherichia, Salmonella, and Vibrio) and many bacteria involved in N fixation. The influence of the nowrecognized Proteobacteria group on rumen metabolism is not known. The rumen microbiome represents a huge resource with great potential. Given the current political climate and public perception in regards to genetically modified organisms searches for natural microorganisms that are capable of toxin degradation and efficient fiber degradation are of utmost importance. Guan et al. (2008) reported that steers with efficient residual feed intake had similar DGGE clusters when compared with inefficient steers. Efficient steers also had nearly twice the concentration of total VFA in their ruminal contents. The next step with this research would be to apply metagenomic sequencing or RNA analysis to determine active metabolic genes and greater microbial community detail. Emphasis also should be placed on studying microbial communities not just individual isolates. Degradation of pyrolizidine alkaloids common in tansy ragwort (Senecio jacobaea) was found to be performed by a consortium of at least 6 microbes isolated from the rumen of sheep (LodgeIvey et al., 2005). Individual isolates of the pyrolizidine alkaloid degrading consortium were not able to degrade the pyrolizidine alkaloids found in tansy ragwort. However, there is evidence in the literature that shows individual bacteria can be responsible for detoxification of toxic plants. Synergistes jonesii is able to detoxify dihydroxypyridine, a toxin found in leucaena (Allison et al., 1992). More recently, a bacterium related to Synergistes jonesii was isolated from the cattle rumen that degraded monofluoroacetate (McSweeney et al., 2009). Monofluoroacetate is a plant toxicant found in 40 species of plants around the world and is responsible for poisoning thousands of livestock (McCosker, 1989). Mining genes of unique rumen bacteria should provide information into their nutritional and physiological requirements making them good candidates for management programs to optimize their detoxifying capacity in the rumen. This may become more important as interest increases in using ruminants to control and minimize invasion of toxic or opportunistic plants that alter native communities as an alternative to costly management methods, such as fire, herbicides, or mechanical clearing. Monitoring and isolating of uncultured bacterial strains from the rumen may be of interest to researcher dealing with grazing ruminants. Scenarios of interest may include observed tolerance to toxic plants, greater production efficiencies or other unique characteristics in populations of domesticated or wild ruminants. One of the main drawbacks to employing nucleic acid-based techniques is that if a bacterium or consortium of bacteria of interest is identified, the bacteria are not isolated using culture independent techniques. However, all is not lost because by classifying bacteria into phylogenetic groups may provide insight into nutritional and physiological requirements of CONCLUSIONS The future of rumen microbiology research is dependent upon the application of molecular research technologies. There is a need to apply modern technologies to improve production efficiency of grazing ruminants. It is important, however, not to become so caught up in the business of surveying rumen microbial communities that the real purpose of the research is forgotten. Focusing attention on how application of modern approaches in rumen microbiology can lead to lowering production input requirements and (or) increasing output of ruminant animal products will augment sustainability of production systems that rely on rangeland resources. LITERATURE CITED Allison, M. J., W. R. Mayberry, C. S. McSweeney, and D. A. Stahl. 1992. Synergistes jonesii, gen. Nov., sp.. nov.: A 27 Li, M., G. G. Penner, E. Hernanadez-Sanabria, M. Oba, and L. L. Guan. 2009. Effects of sampling location and time, and host animal on assessment of bacterial diversity and fermentation parameters in the bovine rumen. J. Appl. Microbiol. 107:1924-1934. Lodge-Ivey. S. L., J. Browne-Silva, and M. B. Horvath. 2009. Technical note: Bacterial diversity in rumen fluid samples collected via oral lavage or rumen cannula. J. Anim. Sci. 87: 2333-2337. Lodge-Ivey, S. L., M. S. Rappe, W. H. Johnston, R. E. Bohlken, and A. M. Craig. 2005. Molecular analysis of a consortium of ruminal microbes that detoxify pyrrolizidine alkaloids. Can. J. Microbiol. 51:455-465. McCosker, T. 1989. Ruminal detoxification of fluoroacetate. Agric. Sci. (New Series) 2:46-47. McSweeney, C., S. Kang, E. Gagen, C. Davis, M. Morrison, and S. Denman. 2009. Recent developments in nucleic acid based techniques for use in rumen manipulation. R. Bras. Zootec. 38:341-351 (supl. especial). Miller, A.J., D.B. Faulkner, R.K. Knipe, D.R. Strohben, D.F. Parrett and L.L. Berger. 2001. Critical control points for profitability in the cow-calf enterprise. Prof. Anim. Sci. 17:295-302. Muyzer, G., E. C. de Waal, and A. G. Uitterlinden. 1993. Profiling of complex microbial populations by denaturing gradient gel electrophoresis analysis of polymerase chain reaction-amplified genes coding for 16S rRNA. Appl. Environ. Microbiol. 59:695–700. Regensbogenova, M., P. Pristas, S. Javorshy, Y. Moon-van der Staay, B. W. van der Staay, H. H. Hackstein, C. J. Newbold, and N. R. McEwan. 2004. Assessment of ciliates in the sheep rumen by DGGE. Lett. Appl. Microbiol. 39:144–147. Rinehart, L. 2008. Ruminant nutrition for graziers. http://attra.ncat.org/attra-pub/PDF/ruminant.pdf. Accessed March 2010. Sundset, M. A., K. I Praestend, I. K. O. Cann, S. D. Mathiesen, and R. I. Mackie. 2007. Novel rumen bacterial diversity in two geographically separated subspecies of reindeer. Microbial. Ecol. 54:424-438. Sundset, M. A., J. E. Edwards, Y. F. Cheng, R. S. Senosiain, M. N. Fraile, K. S. Northwood, K. E. Praesteng, T. Glad, S. D. Mathiesen, and A. G. Wright. 2009. Molecular diversity of the rumen microbiome of Norwegian reindeer on natural summer pasture. 2009. Host Microbe Interactions 57:335-348. Tajima, K., R. I. Aminov, T. Nagamine, H. Matsui, M. Nakamura, and Y. Benno. 2001. Diet-dependent shifts in the bacterial population of the rumen revealed with real-time PCR. Appl. Environ. Microbiol. 67:2766– 2774. Wagner, R. 1994. The regulation of ribosomal RNA synthesis and bacterial cell growth. Arch. Microbiol. 161:100–109. Yu, Z. and M. Morrison. 2004. Comparisons of different hypervariable regions of rrs genes for use in fingerprinting of microbial communities by PCRdenaturing gradient gel electrophoresis. Appl. Environ. Microbiol. 70:4800–4806. rumen bacterium that degrades toxic pyridinediols. Syst. Appl. Microbiol. 15:522-529. Brulc, J. M., C. J. Yeoman, K. E. Nelson, and B. A. White. 2010. Emerging methods in rumen microbiology. Pages 10–21 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Bryant, M. P. and L. A. Burkey. 1953. Cultural methods and some characteristics of some of the more numerous groups of bacteria in the bovine rumen. J. Dairy Sci. 36:205–217. Craig, W. M., G. A. Broderick, and D. B. Ricker. 1987. Quantification of microorganisms associated with the particulate phase of ruminal ingesta. J. Nutr. 117:56–62. Edwards, J. E., A. H. Kingston-Smith, and H. R. Jimenez. 2008. Dynamics of initial colonization of nonconserved perennial ryegrass by anaerobic fungi in the bovine rumen. FEMS Microbial Ecol. 63:141-142. FAO: Food and Agricultrue Organization of the United Nations. 2009. FAO Statistics. http://faostat.fao.org. Accessed October, 2009. Forsberg, C. W. and K. Lain. 1977. Use of adenosine 5triphosphate as an indicator of the microbiota biomass in ramen contents. Appl. Environ. Microbiol. 33:528534. Guan, L. L., J. D. Nkrumah, J. A. Basarab, and S. S. Moore. 2008. Linkage of microbial ecology to phenotype: correlation of rumen microbial ecology to cattle’s feed efficiency. FEMS Microbiol. Lett. 288:85-91. Kang, S., S. E. Denman, M. Morrison, Z. Yu, C. S. McSweeney. 2009. An efficient RNA extraction method for estimating gut microbial diversity by polymerase chain reaction. Curr. Microbiol. 58:464–471 Kocherginskaya, S. A., R. I. Aminov, and B. A. White. 2001. Analysis of the rumen bacterial diversity under two different diet conditions using denaturing gradient gel electrophoresis, random sequencing, and statistical ecology approaches. Anaerobe 7:119–134. Koike, S., J. Pan, Y. Kobayashi, and K. Tanaka. 2003. Kinetics of In Sacco Fiber-Attachment of Representative Ruminal Cellulolytic Bacteria Monitored by Competitive PCR. J. Dairy Sci. 86:1429-1435. Koike, S., Y. Handa, H. Goto, K. Sakai, E. Miyagawa, H. Matsui, S. Ito, and Y. Kobayashi. 2010. Molecular monitoring and isolation of previously uncultured bacterial strains from the sheep rumen. Appl. Environ. Microbiol. 76:1887-1894. Krause, D. O., S. E. Denman, R. I. Mackie, M. Morrison, A. L. Rae, G. T. Attwood, and C. S. McSweeney. 2003. Opportunities to improve fiber digestion in the rumen: microbiology, ecology, and genomics. FEMS Microbiol. Rev. 27:663-693. Kunin, V., A. Copeland, and A. Lapidus et al. 2008 . A Bioinformatician guide to metagenomics. Microbiol. and Mol. Biol. Rev. 72:557-578. Leedle, J.A.Z., M. P. Bryant, and R. B. Hespell. 1982. Diurnal variations in bacterial numbers and fluid parameters in ruminal contents from animals fed low or high forage diets. Appl. Environ. Microbiol. 44:402. 28 PRACTICAL APPLICATION OF MODERN RUMEN MICROBIOLOGICAL TECHNIQUES TO GRAZING RUMINANTS S. L. Lodge-Ivey Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 KEYNOTE: ASSESSMENT OF INTAKE AND DIET COMPOSITION OF GRAZING LIVESTOCK H. Dove1 CSIRO Sustainable Agriculture National Research Flagship, G.P.O. Box 1600, Canberra, A.C.T. 2601, Australia (or at least a sample of the ‘consumed diet’) together with an estimate of total intake. In other words, what do grazing livestock eat and how much? Unfortunately, measuring these under field conditions has always been difficult, mainly due to shortcomings in the available methods of estimation. Additionally, a number of methods can themselves disturb normal grazing behavior and thus the very things they are attempting to measure. This can be a major issue in rangeland environments. Techniques for estimating diet composition and intake have been extensively reviewed (Langlands, 1987; Dove and Mayes, 1996; Mayes and Dove, 2000; Dove and Mayes, 2005). Most were developed with ruminant livestock grazing relatively simple swards and their application to animals in heterogeneous environments introduces further problems. As a generalization, measurements can be based either on the plant biomass or on animal-based measurements. The former, being based on estimates of plant biomass before and after grazing, are difficult to apply under continuous grazing and have the added disadvantage of not providing estimates of intake for individual animals. Sward-based measurements will thus not be discussed further. Rather, this paper will concentrate on selected animal-based techniques, especially markerbased techniques, used to estimate diet composition and intake and will highlight some of the problems associated with their use. The use of the words ‘diet composition’ is deliberate because all these methods measure the outcome (diet composition) rather than the process (diet selection), though they may shed some light on the latter. ABSTRACT: The use of many methods, but particularly methods based on plant-wax markers, for estimating forage intake, diet composition, and supplement intake in grazingbrowsing livestock is reviewed and discussed. The saturated hydrocarbons (alkanes) of plant waxes are shown to be well validated as markers for estimating intake. Alkane-based estimates of intake have the advantage of accommodating individual differences in digestibility and those arising from supplement-forage interactions. Alkanes have also been well validated for estimating diet composition and hence the inputs required to estimate the intake of component species. Further work is still needed on the factors that influence fecal alkane recovery. It is probable that the alkane method can discriminate fewer species in the diet of grazing livestock than they encounter in complex plant communities. A range of approaches for coping with this problem is discussed, such as grouping species in diet composition estimates, combining plant-wax methods, as well as other techniques and the use of other plant-wax markers as diet composition markers. Recent data show that within the other components of plant wax, long-chain alcohols and long-chain fatty acids are very useful additional diet composition markers, which should allow discrimination of a greater number of species in the diet. Supplementary feeds fed to animals can be considered another ‘species’ in the diet and the estimation of supplement intake using plant hydrocarbons, as a special case of the estimation of diet composition, is discussed. Finally, in situations when supplement intakes are known, alkane-based estimates of the supplement proportion in the diet can be used to estimate the intake of all dietary components, without the need to dose animals with synthetic alkanes. ESTIMATION OF INTAKE Under field conditions, the ultimate goal would be to estimate total forage intake and within that, to obtain some estimate whole-diet digestibility. This would allow comparison of nutrient supply with published nutrient requirements (Dove et al., 2010). In an ecological sense, it would also be useful to have an estimate of the intake of forage components (plant species or plant parts) to establish which components of the plant biomass are under the greatest grazing pressure. Additionally, if grazing livestock are also offered supplements, an estimate of supplement intake would be invaluable for defining nutrient supply and quantifying the interaction between supplement and herbage intake. Over the last 60 years, the most successful and widely used methods for estimating intake by grazing livestock have taken advantage of the relationship between intake (I), Key words: intake, diet composition, ruminants, alkanes, plant wax INTRODUCTION Grazing livestock usually consume a diet that differs in terms of plant species, plant parts, and nutrient content from the average of the available plant biomass. It follows that their diet cannot be well quantified using plant-based measurements alone. An accurate estimate of the nutrient intake by grazing animals is more readily made if one has an estimate of both the botanical composition of their diet 1 Corresponding author: [email protected] 31 the digestibility of the whole diet (D) and the resultant fecal output (F), that is: (1) I = F/(1-D) With this general approach, note that intake is estimated from separate estimates of fecal output and digestibility of the consumed diet. There are errors attendant to the estimation of both F and D, but the latter measurement is the greater cause for concern (see below and also Langlands, 1987; Dove and Mayes, 1996; Mayes and Dove, 2000). when diet digestibility is high because the denominator of the equation is so small (Langlands, 1987; Dove and Mayes, 1996; Mayes and Dove, 2000). Esophageal fistulated animals have often been used to obtain samples ‘representing’ the diet of the test animals, although there are increasing concerns about the preparation, maintenance and use of such animals. In vitro estimates of the digestibility of these samples can then be conducted (Tilley and Terry, 1963; Jones and Hayward, 1975). These techniques must be calibrated against in vivo estimates of digestibility. The possible problems associated with this approach have been discussed in detail elsewhere (Langlands, 1987; Dove et al., 2000; Mayes and Dove, 2000). A major issue is the assumption that the samples obtained by fistulated animals represent the diet of intact, test animals; there is no absolute way to test this assumption (Mayes and Dove, 2000). There are doubtless occasions when the assumption is valid (Dove et al., 2000), but this assumption is frequently questioned because esophageal extrusa samples are usually collected over a time-span of minutes, whereas the test animals may be grazing the area for days or weeks. Moreover, the diet selected by the esophageal-fistulated animals may differ from that of the test animals because the former animals are surgicallyprepared, are of different sex or physiological state, or are handled and managed differently from the test animals. Raats et al. (1996) suggested that the possible problems associated with the short collection period used to obtain esophageal extrusa could be overcome by the use of a cannula fitted with a remote-control valve, which in foraging fistulated goats permitted the collection of a diet sample over the course of a whole day, without interrupting normal grazing behavior. A further problem is that the single in vitro estimate of digestibility is applied to all the test animals, despite the strong likelihood that digestibility of a given diet differs between individual animals (Dove et al., 2000) and that these animals may differ markedly in sex, stage of growth, reproductive status, intake level, and even species from the animals used in the original in vitro-in vivo calibration. As will be discussed below, one of the major advantages of estimates of intake based on plant-wax components (e.g., alkanes) is that, like all estimates based on internal markers, differences in digestibility between individuals can be accommodated more adequately, especially if intake is estimated following an estimate of diet composition in individuals. An added complication is that grazing animals are often fed supplements when pasture quantity or quality is inadequate. The in vitro methods cannot accommodate possible interactions between forage and supplement during digestion; for example, the possible reduction in the digestibility of roughages when animals are also consuming starchy components such as cereal grain (Dixon and Stockdale, 1999). It is important to realize that these are potential rather than inevitable sources of error in estimating intake and that accurate data of great scientific and practical importance have been obtained using the Cr-in vitro procedure. Of the problems listed, obtaining a representative sample of the Estimating Fecal Output Under field conditions, total fecal collection is difficult, laborious, and can disturb normal foraging behavior. Fecal output is therefore more commonly estimated from the dilution in feces of a so-called ‘external’ marker, administered orally. A wide range of external markers has been used for the estimation of fecal output (see Table 2 in Mayes and Dove, 2000) but to date, none has proved ideal. Of these, Cr2O3 has been the most widely used and on many (though not all) occasions has performed satisfactorily as a fecal output marker (see examples in Langlands, 1987; Mayes and Dove, 2000). It has been administered orally in gelatin capsules, paper pellets, or by dosing with an intraruminal, controlled-release device (CRD; Luginbuhl et al., 1994). Once fecal Cr concentrations have equilibrated, fecal output is computed from Cr dilution in feces. Possible errors in the use of Cr2O3 for fecal output estimation have been reviewed by Langlands (1987), Parker et al. (1990), Dove and Mayes (1991) and Dove et al. (2000). The use of a CRD to administer Cr helped to overcome some of the problems associated with cyclic changes in fecal Cr concentration arising from once- or twice-daily dosing. However, concerns about possible carcinogenic effects of Cr2O3 have resulted in a major reduction in its use and the Cr-CRD is no longer commercially available. Even if a perfect fecal output marker were available, allowing fecal output to be estimated with little error, a greater cause for concern in the estimation of intake using equation (1) is with the estimate of digestibility. Estimating Whole-Diet Digestibility A major difficulty in obtaining an accurate estimate of whole-diet digestibility is that of obtaining a sample representative of the diet actually consumed by the grazing animal. Even if this is obtained, errors in the actual estimate of digestibility of that sample can seriously reduce the accuracy of the estimate of intake derived using equation (1). Since fecal output is the numerator of this equation, errors of ± 5% in the estimate of fecal output will result in equivalent errors in the estimate of intake (Figure 1). However, because the digestibility term is in the denominator, for a diet of 75% digestibility, errors of ± 5 percentage units in the digestibility estimate will result in errors between -16.7% and +25% in the estimate of intake. If diet digestibility is 85% as in some lush pastures and vegetative crops, the error range encompasses -25% to +50%. Errors in digestibility are thus particularly serious 32 consumed diet is probably the greatest cause for concern because it is the one over which the researcher has least control. A possible advantage of the use of plant alkanes as intake markers is that they can provide a check of the assumption that fistulated and test animals are consuming the same thing (see below and Dove et al., 2000). To avoid some of the problems associated with in vitro estimates of digestibility, an alternative approach is to use indigestible diet components as indicators or ‘internal markers’ of digestibility. In individual animals, the increase in marker concentration between diet and feces provides an estimate of diet indigestibility or the proportion of consumed diet that is excreted in feces, calculated as (marker concentration in feed)/(marker concentration in feces). Digestibility is then calculated as (1 – indigestibility). An advantage of the internal marker approach is that it can be applied to animals consuming a mixture of forage and supplement and will reflect the interactions occurring between these components during digestion. This approach requires that the intake and internal marker concentration of the supplement be known. Mayes and Dove (2000) reviewed the range of dietary components that have been used as internal markers (see their Table 3). The approach has had variable success. In particular, many of the proposed markers are not chemically discrete compounds with specific assays (e.g., ‘chromogen’, potentially indigestible cellulose). This leads to uncertainty about whether the marker assayed in feces is chemically identical to that assayed in the feed and has probably contributed to the variable results used with these empirical internal markers. The unsaturated aliphatic hydrocarbons (alkenes) are relatively common in the floral parts of plants and also in the leaves of a number of tree and shrub species. Most of the alkenes are odd-chain monoenes, with chain lengths ranging from 23 to 33 carbon atoms. Alkenes have potential as diet composition markers (see below), despite their relatively low fecal recoveries. This aspect is discussed in more detail below in relation to the estimation of diet composition. Wax Esters and Free Fatty Acids and Alcohols In most plants, wax esters of saturated, unbranched LCFA and LCOH are the main component of plant wax, with chain lengths in the range C32 to C64. Although both LCOH and, more recently, LCFA have been considered as diet composition markers (Ali et al., 2004; Ali et al., 2005a,b; Fraser et al., 2006; Dove and Charmley, 2008; Lin et al., 2008; Ferreira et al., 2009a), these assessments have been based upon their total concentrations (free plus esterified) because the analytical methods involve a saponification step that cleaves the wax esters (Dove and Mayes, 2006). As yet, neither the intact wax esters nor the free LCFA or LCOH of plant wax have been evaluated as markers in their own right. The LCOH of plant wax are predominantly straightchain, saturated compounds with even-numbered chain lengths up to C34 (Table 1), though in some gymnosperms, LCOH such as 10-C29-alcohol are major components (Table 1) and can function as diet composition markers. As a rule, 1-C30-alcohol is found in much greater concentrations in legumes than grasses (Table 1), and thus can be a useful marker for distinguishing between them. A further useful feature of the LCOH is that in some major pasture grass species with very low alkane concentrations (e.g., Phalaris aquatica), there are abundant quantities of LCOH. The LCFA of plant wax are also mainly straight-chain, saturated compounds with even-numbered chain lengths up to C34 (Table 2). In nutritional studies, LCFA with chain lengths C20 to C34 have been used because they have high fecal recoveries (Grace and Body, 1981; Ali et al., 2004; Ferreira et al., 2009a) and because, unlike some shorter LCFA, they derive exclusively from plant material. The wide variation in LCFA and LCOH patterns in different plants (Tables 1, 2) combined with their relatively high fecal recoveries makes LCFA and LCOH potentially useful diet composition markers (see below). There is also potential to use them as intake markers. USING PLANT CUTICULAR WAX MARKERS IN STUDIES OF THE INTAKE OF HERBIVORES The Nature of Plant Cuticular Wax The cuticular wax on the external surface of plants is a complex mixture of aliphatic lipid compounds, the chemical composition of which differs greatly among different plant species and, to a lesser extent, between different parts of the plant (Dove et al., 1996, 1999; Piasentier et al., 2000; Smith et al., 2001; Bugalho et al., 2004; Ali et al., 2005a,b; Lin et al., 2008). The composition of plant waxes has been comprehensively reviewed elsewhere (Dove and Mayes, 1991, 2005, 2006) and this paper will only discuss the main plant wax compounds that have been used in or offer potential for studies of herbivore nutrition. Principally, this means aliphatic hydrocarbons, long-chain alcohols (LCOH) and long-chain fatty acids (LCFA). Using n-alkanes to Estimate Intake Hydrocarbons The saturated hydrocarbons were initially suggested as digestibility markers (Mayes and Lamb, 1984), but are not ideal for this purpose because they have fecal recoveries of less than 100%. The conceptual leap that allowed their use as intake markers was the realization by Mayes et al. (1986) that if animals are dosed orally with a synthetic even-chain alkane of similar recovery to a plant alkane of adjacent carbon-chain length, then the fecal recovery of the markers is no longer an issue. In essence, the plant alkane is Aliphatic hydrocarbons are present in the waxes of most higher plants, though only rarely as the main component; nalkanes are the most common hydrocarbons. They are present as mixtures with chain lengths ranging from 21 to 37 carbon atoms, with over 90% by weight of the n-alkanes being those with odd-numbers of carbon atoms; C29, C31 and C33 alkanes predominate in most forage or browse species (Table 1). 33 LCOH coupled with either C32 or C36 alkane as the dosed marker. The alkane method for estimating intake offers a number of advantages over other techniques: 1. Provided that the sample of vegetation analyzed for alkanes is representative of the actual diet consumed, the method gives intake estimates that can be regarded as ‘individual’, to the extent that they accommodate the digestibility of the diet in individual animals. Intake estimates are even more truly ‘individual’ if the composition of the diet consumed by individual animals can also be computed using alkanes (see below). Regardless of whether alkane dose rates from an intra-ruminal alkane CRD are measured directly in a sub-group of animals, or assumed from the manufacturer’s specifications, note that the application of a single alkane dose rate to all animals (e.g., when using an alkane CRD) renders the intake estimates less independent and thus less individual. 2. Dosing with an even-chain alkane such as C32 alkane allows the estimation of intake, but does not itself allow an estimate of whole-diet digestibility. If the animals are also dosed with an external marker, fecal output could be estimated and from this, whole-diet digestibility. The marker Cr2O3 could be used but hexatriacontane (C36 alkane) can be used instead (Charmley and Dove, 2007; Dove and Charmley, 2008), which is the reason for its inclusion in the alkane CRD. This permits the estimation of fecal output without having to conduct a separate analysis for fecal Cr. 3. The method can be extended to the simultaneous estimation of diet composition so that the intake of individual plant species, plant parts or both can be estimated (see below). 4. It can be used with animals that are also receiving supplementary feeds, provided the alkane concentrations of the supplement and supplement intake are known. Methods for estimating supplement intake are discussed below. 5. The plant, supplement, and fecal marker alkane concentrations are determined at the same time by the same analytical procedure, which reduces analytical error and bias. Because it is the ratio of the concentrations of alkanes in feces which is used, it is in fact not even necessary to obtain absolute fecal concentrations. 6. If the assumption of equal fecal recoveries of the alkane pair is contravened, in comparative terms, the error in estimated intake that arises is much less than that associated with an equivalent error in the digestibility estimate in equation (1) and more comparable with that associated with errors in fecal output (Figure 1). It can be calculated from equation (2) that if the fecal recoveries of the alkane pair used to estimate intake do differ, each 1% difference in recovery will result in a 1.1% error in the estimate of intake, as shown in Figure 1. The error in intake would thus be expected to be: closely related to the differences in alkane fecal recoveries; much less than that associated with errors in the digestibility estimate, and; comparable with the intake errors arising from errors in the fecal output estimate. functioning as an internal marker to provide an estimate of digestibility, while the dosed alkane functions as the external, fecal output marker. However, in practice and in contrast with equation (1) above, digestibility and fecal output are not calculated separately but rather, for a given alkane pair (plant odd-chain alkane i and dosed even-chain alkane j), intake is calculated directly from the following equation, which can be derived from equation (1) (Dove and Mayes, 1991): (2) Dose ratej Intake = Fecal contentj * Recoveryi * Herbage conenti - Herbage contentj Fecal contenti * Recoveryj There are several key points to note about equation (2): 1. It is ultimately the ratio of the fecal alkane concentrations that is important, not their absolute concentrations. This is of particular relevance to the issue of possible diurnal variation arising from alkane dosing, in that there can be diurnal variation in the concentration of alkanei or of alkanej, and this diurnal variation could be a problem if either alkane concentration was used on its own. However, as Dove et al. (2002) demonstrated, there can be diurnal variation in alkane concentrations but no diurnal variation in the ratio of these concentrations, because their diurnal variation is similar. This renders the alkane method less susceptible to temporal changes in alkane concentrations. 2. If the fecal recoveries of the plant alkane (recoveryi) and the dosed alkane (recoveryj) are equal, they cancel out in equation (2). As a result, errors associated with incomplete fecal recovery also cancel out and an unbiased estimate of intake will be obtained. This is in contrast to methods based on the separate use of internal and external markers to estimate digestibility and fecal output separately, where errors arising from the incomplete recoveries of each marker are additive. 3. It does not matter if the herbage also contains the dosed alkane; this is allowed for explicitly in the denominator of equation (2). The work of Mayes et al. (1986) and many subsequent studies (Dove and Oliván, 1998; Charmley and Dove, 2007; Oliván et al., 2007; Elwert et al., 2008) demonstrated that, in ruminant livestock, the fecal recovery of alkanes increases in curvilinear fashion with carbon-chain length and approaches complete recovery for alkanes C33 and above (see discussion below under ESTIMATION OF DIET COMPOSITION). As a result, adjacent alkanes such as C32 and either C31 or C33 would be expected to have incomplete but similar fecal recoveries. This means that intake can be estimated accurately using, for example, dosed C32 alkane and either C31 or C33 alkane from the plant. It is useful to note in passing that equation (2) does NOT require that each of the two markers be alkanes. Intake could equally well be computed using a plant LCOH or LCFA, together with a dosed alkane, provided the assumption of similar fecal recovery of dosed and natural marker was met. Dove and Charmley (2008), for example, achieved satisfactory estimates of intake using natural C28 34 to establish the exact release rate under the specific conditions of the experiment (Ferreira et al., 2004). These can be obtained either by monitoring CRD release in rumen fistulated animals (Dove et al., 2002; Ferreira et al., 2004; Oliveira et al., 2008) or by taking rectal grab samples beyond the pay-out period of the CRD, to establish the time when alkane release ceases (Charmley and Dove, 2007). Despite occasional problems with their use (Charmley et al., 2003), the alkane CRD greatly reduced the labor requirement for estimating intake under field conditions. Unfortunately, a commercial decision was made in early 2008 to stop producing the devices so that users of the alkane technique now have to employ one of the other dosing procedures described above or to devise another means of administering the markers. One alternative approach is to incorporate the markers into a supplement; this approach is described in more detail below. Oliván et al. (2007) and Charmley and Dove (2007) tested this proposition by estimating intake in housed animals from a range of alkane pairs, some of which differed markedly in fecal recovery. Their recalculated data are shown in Figure 2. Errors in estimated intake were linearly related to the observed differences in fecal recovery of the alkanes used in the estimate by an equation that did not differ significantly from that expected on theoretical grounds from equation (2). Indoor validation studies have shown that the alkane procedure provides reliable estimates of measured intake in a range of animal species (Table 3) and across a wide range of forage types. By definition, validation under field conditions is not possible, because ‘actual’ intakes are unknown and because alternative techniques for estimating intake may be no better than and possibly inferior to the alkane method with which they are being compared. In grazing ewes, Dove et al. (2000) found that there was no simple relationship between estimates of herbage intake based on the alkane or Cr-in vitro methods. The relationship between the two methods for estimating herbage intake was affected both by the physiological state of the animals (late pregnancy v. early lactation v. mid-lactation) and by the level of intake. The major cause of the discrepancy between the intakes estimated by the two different methods was that in vitro estimates of digestibility did not adequately represent in vivo digestibility. Their results indicate that in contrast to the Cr-in vitro procedure, the alkane procedure accommodated the real differences in digestibility between animals. A major consideration in using the alkane method is that the procedures used for dosing of the even-chain alkane do not generate cyclic temporal variation in the fecal alkane ratio used in the intake calculation. Even-chain alkanes have now been administered to sheep or goats using a wide range of carrier matrices, including shredded paper (Mayes et al., 1986; Valiente et al., 2003; Giraldez et al., 2004; Oliván et al., 2007), cellulose powder in gelatin capsules (Dove et al., 2000; Giraldez et al., 2004), paper bungs or filter tips (Lewis et al., 2003; Giraldez et al., 2004), xanthan gum suspension (Marais et al., 1996; Mann and Stewart 2003), or by spraying onto or mixing into dietary components (Unal and Galsworthy, 1999; Elwert and Dove, 2005; Charmley and Dove 2007). The advantages and disadvantages of most of these approaches have been discussed by Dove and Mayes (2006). An intra-ruminal alkane CRD was also developed, tested (Molle et al., 1998; Dove et al., 2002; Ferreira et al., 2004), and commercially released. In addition to containing C32 alkane for the estimation of intake, the commercial alkane CRD also contained C36 alkane as a fecal output marker, which also allowed for an estimate of in vivo digestibility of the diet consumed by the grazing animal (Decandia et al., 2000). These devices greatly reduced labor inputs by delivering alkane at a known rate over a period of 20 d after insertion into the rumen; fecal sampling commenced 5 to 7 d after insertion. They have been shown to give accurate estimates of intake in sheep and cattle (see Table 3 and Molle et al., 1998; Dove et al., 2002; Ferreira et al., 2004). Under grazing conditions, the alkane CRD proved to be a convenient mode of alkane delivery, but care was needed ESTIMATION OF DIET COMPOSITION Animal-based methods for estimating the species composition of the diet have included: microhistological examination of plant fragments (usually cuticle) in esophageal extrusa, a gut compartment, or in feces; and the estimation of carbon-isotope ratios in such samples or in wool or hair, and; the use of plant-wax components as markers. The merits of these various methods have been reviewed extensively (Holechek et al., 1982; Dove and Mayes, 1996; Mayes and Dove, 2000) and in the present paper, the first two approaches will be discussed only briefly. Using Esophageal-Fistulated Animals In domestic herbivores at least, esophageal-fistulated animals have been used to estimate the composition of the diet in terms of plant species (Salt et al., 1994; Dove et al. 2000), plant parts (Dove et al., 1999) or both (Dove et al., 2003). Several authors have presented data to indicate that fistulated and test animals have selected similar diets (Dove et al., 2000), but this cannot always be assumed to be the case. Before the use of alkanes as diet composition markers, there was no way to test this assumption. However, the use of alkanes to estimate diet composition now permits the argument that if fistulated and test animals are consuming the same diet, then their fecal alkane profiles should be similar, or there should be similarity between the alkane profile of esophageal extrusa collected by the fistulated animals and the feces of the test animals (after correction to 100% fecal alkane recovery, see below). Dove et al. (2000) adopted the latter approach to investigate whether esophageal samples collected by castrate male sheep could be assumed to reflect the diet of grazing ewes and demonstrated very close correspondence between the alkane profiles of esophageal samples from fistulated animals and feces samples from either pregnant or lactating ewes. This suggests that the ewes were selecting similar diets. Similarities in the alkane-based estimates of diet composition of fistulated and test animals were also observed by Dove et al. (1999) in sheep grazing senescing pasture. 35 combined with the 13C procedure (García et al., 1999) or with behavioral studies based on active transponder systems (Swain et al., 2008). Microhistological Procedures The microhistological approach is based upon the identification of plant tissue fragments in prepared samples of esophageal extrusa, stomach contents, or feces (Holechek et al., 1982). Diet composition is then estimated in terms of the proportion of these fragments coming from each plant species. In order to relate diet composition and intake, calibrations will usually be needed to convert these proportions to an estimate of diet composition on a DM basis. The method has been used extensively with wild animals, often using stomach contents from slaughtered animals (Holechek et al., 1982). Fecal sampling is more generally applicable and allows repeat sampling and larger numbers of animals to be used, but also has major disadvantages, principally the effect of differential digestion of fragments arising from the different plant species. There is also the problem that a large proportion of fragments can remain unidentifiable, which reduces the quantitative reliability of the method. Nevertheless, the approach can be very useful for establishing the presence or absence of a particular plant species in the diet and can thus complement other methods, such as the use of plant wax markers by ensuring that particular species are included in the estimate of diet composition. Using Plant Wax Markers for Diet Composition Because plant species differ in their cuticular wax profiles of alkanes, LCOH, and LCFA, in theory, any or all of these compounds could be used to estimate diet composition. In practice, the alkanes have been more widely used as diet composition markers (Mayes and Dove, 2000) because they are easier to analyze, but there is increasing use of the LCOH (Bugalho et al., 2004; Dove and Charmley, 2008). A major advantage of using plant wax markers is that diet composition is estimated in the actual test animals so that the resultant alkane concentrations in their consumed diet can be calculated and used as inputs in equation (2) to estimate herbage intake (Mayes et al., 1986). This obviates the need to use esophageal-fistulated animals to obtain a sample of ‘consumed diet’, and on scientific, labor, and animal welfare grounds, this must be seen as a major advantage of the alkane approach. However, it also means that if intake is estimated using the diet composition estimate to provide the ‘herbage’ alkane concentrations, the accuracy of the intake estimate is itself a function of the accuracy of the diet composition estimate. Care must thus be exercised in estimating diet composition, or the estimate of herbage intake may also be incorrect. Stable Isotope Discrimination and Spectroscopic Procedures Tissue from plants that exhibit the C3 or C4 pathways of photosynthesis contains different ratios of the 13C and 12C isotopes of carbon. This difference (δ13C) can be used to estimate the proportion of C3 and C4 herbage in the diet based on esophageal samples (Coates et al., 1987) or fecal samples (Jones et al., 1979). This is especially useful in tropical grazing systems in which grass species make up the main source of C4 plants whereas legumes, browse, and forbs provide most of the C3 plants. Conversely, the method is of limited use in colder environments, where there are few C4 plants. When fecal samples are used, the effect of differential digestion remains an issue and the method will tend to underestimate the intake of plants of high digestibility because less of their carbon remains in feces. Moreover, the fecal carbon isotope ratio can be perturbed by fecal endogenous carbon. The method cannot resolve the diet to the species level, but is still useful in providing an estimate of legume content of the diet, a key variable in many management systems. Spectroscopic procedures, such as laser-induced fluorescence spectroscopy (Anderson et al., 1996) and particularly fecal near infra-red reflectance spectroscopy (F.NIRS; Kronberg and Grove, 2000; Garnsworthy and Unal, 2004; Keli et al., 2008; Dixon and Coates, 2009) also show promise for estimating diet composition. It seems unlikely that the latter technique could be used to quantify a suite of alkanes that differ only by single methyl groups (S.W. Coleman, personal communication). However, there seems to be great scope to combine alkane and spectroscopic approaches for estimating diet composition (Keli et al. 2008). Similarly, alkane-based methods can be Using n-alkanes as Diet Composition Markers There are marked differences in the alkane profiles of plants, both in forage and in browse species (Table 1). These differences have been used to estimate the species composition of esophageal extrusa (Salt et al., 1994; Dove et al., 2000) and have been validated and used as markers for estimating the diet composition of sheep (Mayes et al., 1994; Dove and Mayes, 1996; Kelman et al., 2003; Lewis et al., 2003; Valiente et al., 2003; Lin et al., 2007), cattle (Hameleers and Mayes, 1998; Ferreira et al., 2007a), goats (Merchant, 1996; Brosh et al., 2003), and equines (Ferreira et al., 2007a; Smith et al., 2007). The use of alkanes (or any other plant wax marker) to estimate the composition of a mixture relies on the principle that the alkane profile in the mixture (be it esophageal extrusa, digesta, or feces) must arise from some combination of the alkane profiles in the components that might contribute to the mixture of plant species and (or) plant parts. If the number of marker alkanes at least equals the number of diet components, then in theory at least, diet composition could be calculated using simultaneous equations. However, this approach rapidly becomes computationally difficult with mixtures of 4 or more dietary components. Perhaps more importantly, if there are more alkanes available as markers than there are dietary components, the use of simultaneous equations can involve prior and possibly arbitrary choice of which alkanes to use in the calculations. Moreover, the information provided by those markers that are not included is lost. In such situations, least-squares optimization methods are more 36 satisfactory, in which the best estimate of diet composition is sought using all the data. A number of similar algorithms have been described and used (e.g. Mayes et al., 1994; Dove and Moore, 1995; Newman et al., 1995; Martins et al., 2002; Fraser et al., 2006) and there have been a number of comparisons of these algorithms. For example, Hameleers and Mayes (1998) compared 3 different leastsquares procedures for estimating diet composition using alkanes (Mayes et al., 1994; Dove and Moore, 1995; Newman et al., 1995) and found that they gave almost identical answers. The mathematical procedures available to estimate diet composition are thus satisfactory. It is more important to pay proper attention to fecal recovery corrections and to the choice of which alkanes to use in the calculations. Basically, all these algorithms attempt to minimize the squared deviations between the observed alkane concentrations in feces (or other mixtures), appropriately corrected for incomplete alkane recovery, and the concentration profile arising from the estimate of diet composition. Thus, for a simple example of 3 dietary components: (3) Σ[Actual – Estimated]2alk:1…n fecal alkane recoveries usually show a curvilinear increase with alkane chain length (Mayes et al., 1986; Dove and Oliván, 1998; Moshtagi-Nia and Wittenberg, 2002; Charmley and Dove, 2007; Oliván et al., 2007; Ferreira et al., 2009b). As can be seen from the data in Figure 3(a), published recoveries show a quite consistent pattern, both within and between the major ruminant livestock species (sheep, cattle, and goats). Brosh et al. (2003) reported a linear increase in odd-chain recovery in goats, but when considered in relation to other published recovery data (Figure 3a) it is clear that the general patterns of recovery are similar. There are fewer estimates of fecal alkane recovery in non-ruminant livestock such as equines, but as shown in Figure 3(b), equine fecal alkane recoveries tend to be higher than in ruminants (especially at shorter carbonchain lengths) and much less affected by chain length. As discussed below in relation to alkenes, it may thus be possible to estimate diet selection in equines without fecal recovery corrections because the relative recovery of the different alkanes is so similar. It would be useful to investigate this point further. In correcting fecal alkane concentrations for incomplete recovery, a further assumption made is that the fecal recovery of a given alkane is the same regardless of plant species. Data in support of this assumption remain equivocal. Elwert and Dove (2005), Charmley and Dove (2007), Lin et al. (2007) and Elwert et al. (2008) reported small but significant effects of diet composition on fecal alkane recovery, whereas Ferreira et al. (2007a) noted that as diet digestibility declined, fecal alkane recovery increased. By contrast, other studies have reported that neither diet composition (Brosh et al., 2003; Elwert et al., 2004; Elwert et al., 2006) nor diet digestibility (Elwert et al., 2004) affected fecal alkane recovery. Important as it is, the effect of diet composition or digestibility on fecal alkane recovery thus remains unresolved and would be a fruitful area of further research. The alkane recovery data required to correct fecal concentrations can be obtained from metabolism cage studies with animals fed similar diets. In this case, indoor validation trials have shown that the accuracy of estimated diet composition is best if recovery data for individual animals within a dietary treatment are used, declines when treatment mean recovery data are used, and declines further when recovery data are computed as ‘grand means’ across dietary treatments (Brosh et al., 2003; Elwert et al., 2004; Charmley and Dove, 2007; Ferreira et al., 2007a). As an alternative approach, grazing animals can be dosed with a mix of synthetic alkanes of different chain length (e.g., C24, C28, C32, and C36). The relative recoveries of other alkanes can then be calculated by interpolation (Dove and Mayes, 1996). The relative consistency of the recovery estimates presented in Figure 3(a) suggests that it might be possible to use literature values, though this should be done with caution. In order to provide some quantification of the impact of recovery corrections on estimated diet composition, it is useful to estimate diet composition with and without fecal recovery corrections. Furthermore, an important point to note is that that recovery values can be computed either as absolute recoveries (proportion of known alkane intake or: Σ[Fi – (xAi +yBi + zCi)]2alk:1…n where, for n alkanes: Fi = actual fecal concentration of alkane i x, y, and z = respective amounts of dietary components A, B, and C Ai, Bi, and Ci = respective concentrations of i in A, B, and C. The amounts of dietary components A, B, and C can be converted into dietary proportions of these components from expressions such as: (4) Dietary proportion of component A = x/(x+y+z) As dimensional analysis can demonstrate, x, y, and z in equation (3) can be considered as the amounts of dietary components A, B, and C which, together, will result in 1 kg of feces of the observed concentration profile. It follows that this information can be used to obtain an estimate of whole-diet digestibility, given by: (5) Whole-diet digestibility = ((x+y+z)-1)/(x+y+z) This is not only useful information in its own right, but also serves as a check on whether the least-squares diet composition estimate is ‘realistic’. Fecal Recovery Corrections for Estimating Diet Composition When alkanes are used to estimate herbage intake, fecal alkane concentrations do not require correction for incomplete fecal recovery. By contrast, before alkanes (or any other plant wax marker) can be used as diet composition markers, their fecal concentrations must be corrected for incomplete recovery. In ruminant livestock, 37 obtained if certain alkanes are not used in the calculation (Dove et al., 1999; Lewis et al., 2003; Lin et al., 2007). Theoretically, these would be the alkanes that contribute substantially to the discrepancy between observed and expected concentrations, but little to the actual discrimination of diet components; the issue is how to identify these. This is an issue of current concern and one which requires resolution if plant wax components are to gain wider acceptance as a method for estimating diet composition. Multivariate statistical procedures, such as Canonical Variates Analysis (discriminant analysis) or Principal Components Analysis, can be used to establish whether plant species can be distinguished and which alkanes are best related to the ability to discriminate between diet components (Dove et al., 1999; Piasentier et al., 2000; Bugalho et al., 2004; Charmley and Dove 2007; Lin et al., 2008). It is thus a useful exercise to use such statistical procedures to explore whether it is mathematically likely that the diet components can be discriminated. Conversely, it is important to note that that the demonstration that they can be distinguished in theory does not mean that they will be in practice due to such problems as a given ‘mix’ of some dietary components mimicking the marker profile of another. More dietary components than markers. A more usual situation, especially in species-diverse plant communities such as rangelands, is that there are more plant species available for possible consumption than there are alkanes to discriminate them. Under these circumstances, several general approaches could be attempted in order to obtain a diet composition estimate: 1. Combine the discriminatory power of alkanes and other methods for estimating diet composition. For example, Salt et al. (1994) combined microhistological and alkane procedures to estimate the diet composition of sheep grazing mixed grassland-heathland. Similarly, stable isotope procedures (Jones et al., 1979; Anderson et al., 1996) or F.NIRS (Keli et al., 2008) could be used in combination with alkanes to increase discrimination. In such approaches, one technique could allow a reduction in the number of plant pools to be discriminated by establishing, for example, that particular plant species are not consumed. In this regard, data obtained by close observation of grazing behavior should not be dismissed. 2. Decrease the number of dietary pools to be discriminated by grouping species in the diet on some logical basis or, at the extreme, leaving out species altogether. The latter approach may be valid if there are behavioral or other data indicating the species is never consumed, but it should be used with caution. The grouping of species is another area in which multivariate statistical procedures can be used (Bugalho et al., 2002, 2004; Ferreira et al., 2007b) to indicate that, a priori, it will be difficult to separate some of the available species because they are too similar in their marker profiles. This provides an objective justification for grouping those species. Alternatively, species can be grouped on a ‘functional’ basis such as herbage vs. browse species (Bugalho et al., 2002) or on taxonomic grounds (e.g., grasses vs. clovers vs. Lotus spp.; Kelman et al., 2003). Note that selection by different animals between excreted in feces), or as relative recoveries (the recoveries of a suite of alkanes relative to an assumed value, usually 1, for the longest alkane). For an estimate of diet composition only, relative recoveries are sufficient. This means that under field conditions, animals can be given a known daily dose of a mix of alkanes, including a long-chain alkane (e.g., C36) with high recovery, and the fecal concentrations of dosed alkanes used to get the relative recoveries of other alkanes by interpolation. However, in order to obtain an unbiased estimate of whole-diet digestibility with leastsquares minimization calculations, absolute recovery values must be used. On theoretical grounds, the maximum number of components that can be discriminated in the diet is limited to the number of available alkanes (C21 to C37), thus making it theoretically possible to distinguish more than 15 dietary components. However, in practice fewer dietary components can be discriminated, for two reasons: 1. The odd-chain alkanes usually make up 90% or more of total alkane content and the concentrations of even-chain alkanes are usually much lower (often <10 to 30 mg/kg OM) and may have larger analytical errors, so that they are usually (but not always) less useful as markers. 2. It is likely that the reliability of the diet composition estimate will decline as the number of components increases because there is an increased likelihood that different combinations of components could result in the same fecal alkane patterns. For example, in certain combinations, the alkane profile of a mix of diet components A and B may be indistinguishable from that of component C. Despite the above concerns about factors influencing fecal alkane recovery or how this can be estimated and applied, the results of a number of validation studies indicate that accurate estimates of diet composition can be obtained using this approach. The data in Table 4 are regression relationships between known and estimated diet compositions either published in or recalculated from the results of these studies. In 13 of the 15 examples shown, the relationship between known and estimated diet composition does not differ from the line of equality. Which Alkanes to Use in Diet Composition Calculations? As a generalization, two situations can be considered, the first involving more markers (alkanes) than diet components and the second involving more diet components than there are alkanes to distinguish them. More markers than diet component. In some grazing situations involving simple associations in a renovated pasture planted to improved forage species (e.g., grassclover pastures) there might be more markers (alkanes) available than there are possible species in the diet. As indicated above, the least-squares procedures employed can accommodate information about all available alkanes, but this may not be a sensible approach because it is possible that some alkanes may discriminate much better than others between components. Indeed, other alkanes may prove to be negatively correlated with the capacity to distinguish between dietary components (Dove et al., 1999). In such cases, more reliable estimates of diet composition may be 38 to emphasize the point made above that ultimately it is the relative recoveries of the markers, which are important for estimating diet composition, not the absolute recoveries. In this case, the relative recoveries of alkenes C29 to C33 were so similar that recovery correction did not alter estimated diet composition. Long-chain alcohols. The LCOH are now proving to be useful diet composition markers. Large between-species differences in LCOH concentrations exist (Table 1) and account for 60 to 90% of the variance in LCOH concentrations between plant species (Bugalho et al., 2004; Ali et al., 2005b). In the dicotyledons, concentrations of C30 alcohol are generally much greater than in the monocotyledons, which is likely to be a very useful feature when using alcohol concentrations in diet composition estimates. While primary alcohols predominated in most of the plants analyzed, the secondary alcohol, 10-nonacosanol, could be a potentially important diet composition marker in studies involving gymnosperms. Implicit in the use of LCOH or LCFA profiles together with alkane profiles is the assumption that the first two groups of compounds contribute extra discriminatory information. This assumption needs to be checked because if there are high correlations between the concentration profiles of LCOH, LCFA, and alkanes, it may well be that the LCOH or LCFA provide little extra information. This can be assessed by comparing multivariate statistical analyses of, for example, alkane and LCOH profiles for the same species, using techniques such as Orthogonal Procrustes Rotation. In their estimations of diet composition of housed sheep, Dove and Charmley (2008) used this approach to demonstrate that the information supplied by the LCOH was additional to that supplied by alkanes, a finding confirmed by Lin et al. (2008). As a result, estimates of diet composition based on both markers are better than those based on alkanes alone (Bugalho et al., 2004; Dove and Charmley 2008). The fecal recoveries of dietary LCOH have now been determined in several studies and show a progressive increase with chain length similar to that observed with nalkanes (Figure 4a). In these studies, recoveries have ranged from about 0.6 for C22-OH to about 0.9 for C30-OH and by contrast with alkanes, tend to show a linear increase in fecal recovery with increasing carbon-chain length. The LCOH have now been successfully used in a number of recent estimates of diet composition of livestock (Bugalho et al., 2004; Ali et al., 2005; Fraser et al., 2006; Dove and Charmley, 2008; Lin et al., 2008). Long-chain fatty acids. Procedures for LCFA analysis, as an extension of the procedure used for alkanes and LCOH, have now been published (Ali et al., 2004; Dove and Mayes, 2006). Fecal recoveries of individual LCFA, as with other cuticular wax components, increase with increasing chain length (Figure 4b) in sheep (Ali et al., 2004) and in goats (Ferreira et al., 2009a). The widespread incidence of LCFA in plant cuticular wax, the betweenspecies differences in LCFA profile, and the fecal recovery data to date suggest that these compounds could also be used as diet composition markers. Ali et al. (2005) reported concentration data for the alkanes, LCOH, and LCFA in a range of plant species. The best estimates of diet plant species within such groupings may influence the accuracy of the diet composition estimate. Bugalho et al. (2002) found that after grouping the herbage species into a single component, simulated selection within the grouping only had a minor effect on the estimates of diet composition. However, data of Ferreira et al. (2007b) suggest that this may have been a case-specific outcome and dietary grouping based on multivariate statistical analysis should be used with care. If species are grouped for the purposes of estimating diet composition, it would be sensible to carry out some assessment of the impact of possible selection within the groups. 3. A more recent development is the approach in which diet composition is estimated from ‘possible infinite bounded combinations of plants species’, the upper and lower limits of which are determined by linear programming (Barcia et al., 2007). Those authors suggest that this approach can be applied when the number of markers is lower than the number of species, thus avoiding the need for grouping. Though promising, the approach requires further testing. 4. An obvious approach is to increase the number of ‘discriminators’ used to estimate diet composition; this could be achieved either by combining the use of alkanes (or other plant wax markers) with other methods for estimating diet composition (as described above) or by recruiting further groups of plant wax markers. Using Other Plant Wax Markers to Estimate Diet Composition The use of other plant wax markers to estimate diet composition has been an area of recent activity following the development of reliable analytical procedures for both the LCOH and LCFA (Ali et al., 2004; Dove and Mayes, 2006). At present, the additional components of plant wax that might be considered as diet composition markers are the alkenes (Dove and Oliván, 2006), LCOH (Kelman et al., 2003; Ali et al., 2005b; Fraser et al., 2006; Dove and Charmley, 2008; Lin et al., 2008) and LCFA (Ali et al., 2004; 2005b; Lin et al., 2008; Ferreira et al., 2009a). These components also have different concentrations in the different plant species and plant parts and can be obtained together with alkanes as part of the same analytical procedure (Ali et al., 2004; Dove and Mayes, 2006). Alkenes. The unsaturated hydrocarbons or alkenes tend to be predominant in the floral parts of plants and have also been found in a number of browse species; they are extracted from plant material and feces in the same fraction as n-alkanes. There are only a few estimates of fecal recoveries of plant alkenes, which in sheep are much lower (25 to 40%) than the recoveries of alkanes of equivalent chain length (Dove and Oliván, 2006; Elwert et al., 2004; Elwert and Dove, 2005) and show little variation with carbon-chain length, especially from C29 to C33 (Figure 3(c)). Paradoxically, this means that despite their low fecal recoveries, alkenes C29 to C33 gave excellent estimates of diet composition (Dove and Oliván, 2006), which did not differ from those estimated using alkanes (see below). Moreover, there was no advantage in correcting fecal alkene concentrations for incomplete recovery. This serves 39 In the approaches discussed to this point, to obtain an estimate of total intake and thus of supplement intake, it was necessary to dose the animals with a source of evenchain alkanes (e.g., the alkane CRD). However, recent studies indicate that estimates of forage intake could be obtained without having to separately dose the animals with alkanes. composition were obtained with a combination of all three marker classes (Ali et al., 2005b), indicating that each class of marker provided different discriminatory information. ESTIMATION OF SUPPLEMENT INTAKE In many grazing systems, there are times of the year when the supply or quality of herbage is inadequate for the production target and animals are offered supplementary feed. Interpretation of animal responses and the interaction between herbage and supplement intakes would be much easier if individual intakes of supplement were known. Supplement intake by individual animals has been estimated using markers that are incorporated into and measured in the body water pool of the animal (e.g., tritiated water or lithium salts), or methods involving fecal markers (e.g., chromic oxide, ytterbium salts). The features of these methods have been discussed in detail by Mayes and Dove (2000) and will not be considered further. The present discussion will be restricted to the possible use of plant wax markers for estimating supplement intake, which is in essence a special case of the estimation of diet composition. Using Supplement Intake to Estimate Forage Intake In the above examples, plant wax marker profiles were used to estimate diet composition, which then allowed the calculation of total intake and thus supplement intake in dosed animals. However, there are a number of situations in which the intake of supplement by individual animals either is known or could be known with relatively little trouble; for example, cows receiving known amounts of supplement while in the dairy parlor. If these supplements either contain or can be labeled with alkanes, the proportion of supplement in the diet can be estimated as above. However, because actual supplement intake is also known, total and thus herbage intake can also be computed. For example, if supplement is estimated to comprise 10% of the diet of a sheep and known supplement intake is 150 g/day, then total intake must be 1500 g/day and forage intake 1350 g/day. This approach provides a means of estimating the intake of pasture or its component species, without having to dose animals with alkanes, which in turn, may avoid undue stress or perturbations of normal grazing behavior. Moreover, it may provide an alternative approach to estimating herbage intake, now that the alkane CRD is no longer commercially available. The procedure for estimating intakes in this way is as follows: 1. Estimate diet composition as described above, using the alkane profiles of diet components and of feces (recoverycorrected). In a two-component diet (forage and supplement), if the known supplement intake is Is, and the proportion of supplement in the diet is ps, then the total intake It is given by: (6) It = Is/ps Intake of the forage component of the diet (I f) is then given by (It – Is). 2. This method for two-component diets (forage + supplement) can be restated in more general terms to allow the estimation of the intake of all the forage components of the diet in situations where the animals are consuming known amounts of supplement and unknown amounts of a number of forage species. By rearrangement of the above terms, it can be shown that If can be estimated directly as (7) If = Is*(pf/ps) where pf is the proportion of forage in the diet. The advantage of the approach embodied in equation (7) is its equal applicability to multi-component diets (supplement, forage species 1, forage species 2…forage species n). A further and major advantage is that the intake of the various forage species and (or) plant parts is obtained without the separate dosing of the animals. In effect, all of the alkanes in the supplement are used as the dose. The use of Using Plant Wax Alkanes If the supplement has its own alkane profile or can be labeled with a distinctive profile (Dove and Oliván, 1998, Elwert and Dove 2005; Charmley and Dove, 2007), then the intake of supplement can be estimated by treating it as if it were one of the ‘species’ in the diet. Diet composition is first estimated as described above; this provides an estimate of the proportion of supplement in the diet. If the animals are also dosed with even-chain alkanes (e.g., C32), this allows the estimation of total intake using equation (2) above. Diet composition then allows intake to be partitioned into its components, including supplement, thus providing an estimate of supplement intake. A further advantage of this approach is that it can provide an estimate of whole-diet digestibility, which accommodates, in individual animals, any interactions of supplement and herbage during digestion. If the supplement is itself based on roughage (e.g., hay, silage, or pelletted herbage) then the natural alkanes of the supplement can be employed without further manipulation. However, in some cases (e.g., legume grains, coarse grains, or oilseed meals), the levels of alkane may be insufficient to permit the estimation of diet composition. Under these circumstances, supplements can be labelled with an external source of hydrocarbons, such as beeswax (Dove and Oliván, 1998; Elwert and Dove, 2005). Validation studies (Dove and Oliván, 1998; Figure 5) have shown that this provides accurate estimates of supplement intake. Moreover, as described above, supplement intake estimates were equally good when based on the unsaturated hydrocarbons (alkenes; Dove and Oliván, 2006). It is worth re-emphasizing that if the diet composition estimate is in error, so will be the estimates of supplement and total intake. This means that care must be taken in estimating the diet composition itself. 40 supplement intake to estimate the intake of the forage components of the diet was validated for 2-component diets (Elwert and Dove, 2005; Figure 6a) and for multicomponent diets (Charmley and Dove, 2007; Figure 6b). Regardless of whether individual, treatment mean or grand mean fecal recovery data were used to estimate diet composition, the resultant estimates of intake using this ‘supplement dosing’ approach were close to actual intake (Table 5) and to those estimated from alkane-CRD dosing using either the C32/C33 or C36/C33 alkane pairs (Charmley and Dove 2007; Dove and Charmley, 2008). Ali, H. A. M., R. W. Mayes, B. L. Hector, and E. R. Ørskov. 2005a. Assessment of n-alkanes, long-chain fatty alcohols and long-chain fatty acids as diet composition markers: The concentrations of these compounds in rangeland species from Sudan. Anim. Feed Sci. Tech. 121:257-271. Ali, H. A. M., R W. Mayes, B. L. Hector, A. K. Verma, and E. R. Ørskov. 2005b. The possible use of n-alkanes, long-chain fatty alcohols and long-chain fatty acids as markers in studies of the botanical composition of the diet of free-ranging herbivores. J. Agric. Sci. (Camb) 143:85-95. Anderson, D. M., P. Nachman, R. E. Estell, T. Ruekgauer, K. M. Havstad, E. L. Fredrickson, and L. W. Murray. 1996. The potential of laser-induced fluorescence (LIF) spectra in sheep feces to determine diet botanical composition. Small Rum. Res. 21:1-10. Barcia, P., M. N. Bugalho, M. L. Campagnolo, and J. O. Cerdeira. 2007. Using n-alkanes to estimate diet composition of herbivores: a novel mathematical approach. Anim. 1:141-149. Berry, N. R., M R. L. Scheeder, F. Sutter, T. F.Kröber, and M. Kreuzer. 2000. The accuracy of intake estimation based on the use of alkane controlled-release capsules and faeces grab sampling in cows. Ann. Zootech. 49:313. Brosh, A., Z. Henkin, S. J. Rothman, Y. Aharoni, A. Orlov, and A. Arieli. 2003. Effects of n-alkane recovery in estimates of diet composition. J. Agric. Sci. (Camb) 140:93-100. Bugalho, M. N., R. W. Mayes, and J. A. Milne. 2002. The effects of feeding selectivity on the estimation of diet composition using the n-alkane technique. Grass and Forage Sci. 57:224-231. Bugalho M. N., H. Dove, W. M. Kelman, J. T. Wood, and R. W. Mayes. 2004. Plant wax alkanes and alcohols as herbivore diet composition markers. J. Range Manage. 57:259-268. Charmley, E. and H. Dove. 2007. Using plant wax markers to estimate diet composition and intakes of mixed forages in sheep by feeding a known amount of alkanelabelled supplement. Aust. J. Agric. Res. 58:1215-1225. Charmley, E., D. R. Ouellet, D. M. Veira, R. Michaud, J. L. Duynisveld, and H. V. Petit. 2003. Estimation of intake and digestibility of silage by beef steers using a controlled release capsule of n-alkanes. Can. J. Anim. Sci. 83:761-768. Coates, D. B., P. Schachenmann, and R. L. Jones. 1987. Reliability of extrusa samples collected from steers fistulated at the oesophagus to estimate the diet of resident animals in grazing experiments. Aust. J. Exp. Agric. 27:739-745. Decandia, M., M. Sitzia, A. Cabiddu, D. Kababya, and G. Molle. 2000. The use of polyethylene glycol to reduce the anti-nutritional effects of tannins in goats fed woody species. Small. Rum. Res. 38:157-164. Dixon, R.M. and C. R. Stockdale. 1999. Associative effects between forages and grains: consequences for feed utilisation. Aust. J. Agric. Res. 50:757-773. CONCLUSIONS In a variety of species of grazing livestock, the alkane procedure for estimating forage intake is now well validated. Under field conditions, the procedure has been compared with previous techniques such as the Cr-in vitro procedure. Although no simple relationship between these approaches should be expected, the alkane-based estimates are likely to be more accurate because the natural alkane functions essentially as an internal marker and thus accommodates differences in forage digestibility between individuals and those arising from interactions between supplement and forage. A further (and major) advantage of the use of plant wax markers is that they also allow an estimate of diet composition and thus the partitioning of total intake into its component plant species. Estimates of diet composition require correction for incomplete fecal marker recovery and more work is required to quantify the relative effects of animal species and particularly plant species on the recovery of a given alkane. If estimates of whole-diet digestibility are wanted, actual fecal alkane recoveries must be used, but for an estimate simply of diet composition, relative recoveries (i.e., the recovery of the alkanes relative to each other) will suffice. When used alone, alkanes can discriminate between fewer species than is often encountered by grazing livestock utilizing complex plant communities. To address this, the LCOH and LCFA of plant wax can now also be used as diet composition markers. This, plus the use of multivariate statistical procedures to guide the grouping of species within the possible diet, should extend substantially the number of species that can be discriminated in the diet. In addition, other compounds in plant wax, as yet unevaluated as markers, could further extend diet composition estimates. Supplement intake can also be estimated using wax markers, and there is considerable scope for exploiting known supplement intakes as a means of estimating the intake of other diet components, thus avoiding the need to separately dose the test animals with synthetic alkanes. LITERATURE CITED Ali, H. A. M., R. W. Mayes, C. S. Lamb, A. K. Verma, and E. R. Orskov. 2004. The potential of long-chain fatty alcohols and long-chain fatty acids as diet composition markers: Development of methods for quantitative analysis and fecal recoveries of these compounds in sheep fed mixed diets. J. Agric. Sci. (Camb) 142:71-78. 41 Dixon, R. M. and D. B. Coates. 2009. Near infrared spectroscopy of faeces to evaluate the nutrition and physiology of herbivores. J. Near Infrared Spectroscopy 17:1-31. Dove, H. and E. Charmley. 2008. Using the alkanes and long-chain alcohols of plant cuticular wax to estimate diet composition and the intakes of mixed forages in sheep consuming a known amount of alkane-labelled supplement. Anim. 2:1474-1485. Dove, H. and R. W. Mayes. 1991. The use of plant wax alkanes as marker substances in studies of the nutrition of herbivores: a review. Aust. J. Agric. Res. 42:913-952. Dove, H. and R. W. Mayes. 1996. Plant wax components: A new approach to estimating intake and diet composition in herbivores. J. Nutr. 126:13-26. Dove, H. and R. W. Mayes. 2005. Using n-alkanes and other plant wax components to estimate intake, digestibility and diet composition of grazing/browsing sheep and goats. Small Rum. Res. 59:123-139. Dove, H. and R. W. Mayes. 2006. Protocol for the analysis of n-alkanes and other plant-wax compounds and for their use as markers for quantifying the nutrient supply of large mammalian herbivores. Nature Protocols 1:1680-1697. (http://www.nature.com/nprot/journal/v1/n4/index.html) Dove, H. and A. D. Moore. 1995. Using a least-squares optimisation procedure to estimate botanical composition based on the alkanes of plant cuticular wax. Aust. J. Agric. Res. 46:1535-1544. Dove, H. and M. Oliván. 1998. Using synthetic or beeswax alkanes for estimating supplement intake in sheep. Anim. Prod. Aust. 22:189-192. Dove, H. and M. Oliván. 2006. The possible use of the alkenes (unsaturated hydrocarbons) of plant cuticular wax as diet composition markers in sheep. Page 1-7 In: Herbivores: The Assessment of Intake, Digestibility and the Roles of Secondary Compounds. C.A. SandovalCastro, F.D. DeB Hovell, J.F.J. Torres-Acosta and A. Ayala-Burgos (eds). BSAS Publication 34. ISBN 1904761-52-6. Satellite Meeting of VI Internatl. Symp. Nutr. Herbivores. Mérida, Mexico, October 2003. Dove, H., R. W. Mayes, and M. Freer. 1996. Effects of species, plant part and age on the n-alkane concentrations in the cuticular wax of pasture plants. Aust. J. Agric. Res. 47:1333-1347. Dove, H., P. E. Quigley, and J. Z. Foot 2003. Estimating the proportions of different plant species and plant parts in the diet of ewes grazing grass/clover pastures. Proc. IX World Congr. Anim. Prod., Porto Alegre, Brazil, October 2003. Dove, H., J. T. Wood, R. J. Simpson, B. J. Leury, K. L. Gatford, C. Siever-Kelly, and T. A. Ciavarella. 1999. Spray-topping annual grass pasture with glyphosate to delay loss of feeding value during summer. III. Quantitative basis of the alkane-based procedures for estimating diet selection and herbage intake by grazing sheep. Aust. J. Agric. Res. 50:475-485. Dove, H., M. Freer, and J. Z. Foot. 2000. The nutrition of grazing ewes during pregnancy and lactation: a comparison of alkane-based and chromium/in vitrobased estimates of herbage intake. Aust. J. Agric. Res. 51:765-777. Dove, H., R. W. Mayes, C. S. Lamb, and K. J. Ellis. 2002. Factors influencing the release rate of alkanes from an intra-ruminal, controlled-release device, and the resultant accuracy of intake estimation in sheep. Aust. J. Agric. Res. 53:681-696. Dove, H., S. G. McLennan, and D. P. Poppi. 2010. Application of nutrient requirement schemes to grazing animals. Pages 133–149 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Elwert, C. and H. Dove. 2005. Estimation of roughage intake in sheep using a known daily intake of a labelled supplement. Anim. Sci. 81:47-56. Elwert, C., H. Kluth, and M. Rodehutscord. 2004. Effect of variable intake of alfalfa and wheat on faecal alkane recoveries and estimates of roughage intake in sheep. J. Agric. Sci. (Camb) 142:213-223. Elwert, C., H. Dove, H., and M. Rodehutscord. 2006. Effect of roughage species on faecal alkane recovery in sheep, and the effect of drying treatment on alkane concentration. Aust. J. Exp. Agric. 46:771-776. Elwert, C., H. Dove, and M. Rodehutscord. 2008. Faecal alkane recoveries in multi-component diets and subsequent estimates of diet composition in sheep. Anim. 2:125-134. Ferreira, L. M. M., M. Oliván, M. A. M. Rodrigues, K. Osoro, H. Dove, and A. Dias-da-Silva. 2004. Estimation of feed intake by cattle using controlled-release capsules containing n-alkanes or chromium sesquioxide. J. Agric. Sci. (Camb) 142:225-234. Ferreira, L. M. M., U. Garcia, M. A. M. Rodrigues, R. Celaya, A. Dias-da-Silva, and K. Osoro. 2007a. The application of the n-alkane technique for estimating the composition of diets consumed by equines and cattle feeding on upland vegetation communities. Anim. Feed Sci. Tech. 138:47-60. Ferreira, L. M. M., M. Oliván, R. Celaya, U. Garcia, M. A. M. Rodrigues, and K. Osoro. 2007b. The use of nalkanes to estimate diet composition of ruminants grazing on species diverse plant communities – Effect of feeding selectivity on diet composition estimates. Livest. Sci. 111:114-123. Ferreira, L. M. M., S. Carvalho, V. Falco, R. Celaya, U. García, A. S. Santos, M. A. M. Rodrigues, and K. Osoro. 2009a. Assessment of very long-chain fatty acids as complementary or alternative natural fecal markers to nalkanes for estimating diet composition of goats feeding on mixed diets. J. Anim. Sci. 87:2732-2745. Ferreira, L. M. M., R. Celaya, U. García, M. A. M. Rodrigues, and K. Osoro. 2009b. Differences between domestic herbivores species in alkane faecal recoveries and the accuracy of subsequent estimates of diet composition. Anim. Feed Sci. Tech. 151:128-142. 42 Fraser, M. D., V. J. Theobald, and J. M. Moorby. 2006. Determining diet composition on complex swards using n-alkanes and long-chain fatty alcohols. Ecol. Appl. 16:1901-1910. García, S. C., C. W. Holmes, J. Hodgson, and A. Macdonald. 1999. The combination of the n-alkanes and 13 C methods as a new tool for estimating individual DM intakes of pasture and maize silage by grazing dairy cows. Proc. NZ Soc. Anim. Prod. 59:156-157. Garnsworthy, P. C. and Y. Unal. 2004. Estimation of drymatter intake and digestibility in group-fed dairy cows using near infrared reflectance spectroscopy. Anim. Sci. 79:327-334. Gedir, J. V. and R. J. Hudson. 2000. Estimating dry matter digestibility and intake in wapiti (Cervus elaphus canadensis) using the double n-alkane ratio technique. Small. Rum. Res. 36:57-62. Giraldez, F. J., C. S. Lamb, S. López, and R. W. Mayes. 2004. Effects of carrier matrix and dosing frequency on digestive kinetics of even-chain alkanes and implications on herbage intake and rate of passage studies. J. Sci. Food Agric. 84:1562-1570. Grace, N. D. and D. R. Body. 1981. The possible use of long-chain (C19-C32) fatty acids in herbage as an indigestible faecal marker. J. Agric. Sci. (Camb) 97:743745. Hameleers, A. and R. W. Mayes. 1998. The use of nalkanes to estimate herbage intake and diet composition by dairy cows offered a perennial ryegrass/white clover mixture. Grass Forage Sci. 53:164-169. Holechek, J. L., M. Vavra, and R. D. Pieper. 1982. Botanical composition determination of range herbivore diets: a review. J. Range Manage. 35:309-315. Jones, D. I. H. and M. V. Hayward. 1975. The effect of pepsin pretreatment of herbage on the prediction of dry matter digestibility from solubility in fungal cellulose solution. J.Sci. Food Agric. 26:711-718. Jones, R. J., M. M. Ludlow, J. H. Throughton, and C. G. Blunt. 1979. Estimation of the proportion of C3 and C4 plant species in the diet of animals from the ratio of natural 12C and 13C isotopes in the faeces. J. Agric. Sci. (Camb) 92:91-100. Keli, A., D. Andueza, A. de Vega, and J. A. Guada. 2008. Validation of the n-alkane and NIRS techniques to estimate intake, digestibility and diet composition in sheep fed mixed lucerne:ryegrass diets. Livest. Sci. 119:42-54. Kelman, W. M., M. N. Bugalho, and H. Dove. 2003. Cuticular wax alkanes and alcohols used as markers to estimate the diet composition of sheep (Ovis aries). Biochem. Syst. Ecol. 31:919-927. Kronberg, S. L. and A. V. Grove. 2000. A potential simple and inexpensive method for estimating CP and forage intake by grazing cattle. Proc. West. Sect. Am. Soc. Anim. Sci. 51:320-323. Langlands, J. P. 1987. Assessing the nutrient status of herbivores. Pages 363-390 In: The Nutrition of Herbivores. J.B. Hacker and J.H. Ternouth (eds). Academic Press, Sydney. Lewis, R. M., A. M. Magadlela, N. S. Jessop, and G. C. Emmans. 2003. The ability of the n-alkane technique to estimate intake and diet choice of sheep. Anim. Sci. 77:319-327. Lin, L. -J., H. -L. Luo, Y. -J. Zhang, and B. Shu. 2007. The effects, in sheep, of dietary plant species and animal live weight on the faecal recovery rates of alkanes and the accuracy of intake and diet composition estimates obtained using alkanes as faecal markers. J. Agric. Sci. (Camb) 145:87-94. Lin, L. J., H. L. Luo, H. Wang, Y. J. Zhang, and B. Shu. 2008. Evaluation of long-chain alcohols and fatty acids, in combination with alkanes, as markers in the estimation of the composition of four herbages in mixtures. Grass Forage Sci. 64:19-25. Luginbuhl, J-M., K. R. Pond, J. C. Burns, and D. S. Fisher. 1994. Evaluation of the Captec controlled release chromic oxide capsule for fecal output determination in sheep. J. Anim. Sci. 72:1375-1380. Mann, J. and P. G. Stewart. 2003. Kikuyu (Pennisetum clandestinum) intake determined by alkanes administered in a xantham gum suspension. S. Afr. J. Anim. Sci. 33:27-31. Marais, J. P., D. L. Figenschou, P. L. Escott Watson, and L. N. Webber. 1996. Administration in suspension form of n-alkane external markers for dry matter intake and diet selection studies. J. Agric. Sci. (Camb) 126:207-210. Martins, H., D. A. Elston, R. W. Mayes, and J. A. Milne. 2002. Assessment of the use of n-alkanes as markers to describe the complex diets of herbivores. J. Agric. Sci. (Camb) 138:425-434. Mayes, R. W. and H. Dove. 2000. Measurement of dietary nutrient intake in free-ranging mammalian herbivores. Nutr. Res. Rev. 13:107-138. Mayes, R. W. and C. S. Lamb. 1984. The possible use of nalkanes as indigestible faecal markers. Proc. Nutr. Soc. 43:39A. Mayes, R. W., C. S. Lamb, and P. M. Colgrove. 1986. The use of dosed and herbage n-alkanes as markers for the determination of herbage intake. J. Agric. Sci. (Camb) 107:161-170. Mayes, R. W., N. A. Beresford, C. S. Lamb, C. L. Barnett, B. J. Howard, B-E. V. Jones, O. Eriksson, K. Hove, O. Pedersen, and B. W. Staines. 1994. Novel approaches to the estimation of intake and bioavailability or radiocaesium in ruminants grazing forested areas. Sci. Total Environ. 157:289-300. Merchant, M. 1996. The intake of grass and rush (Juncus effusus L.) by goats grazing rush-infested pasture. Grass Forage Sci. 51:81-87. Molle, G., M. Decandia, and H. Dove. 1998. A comparison between different procedures for dosing n-alkanes to sheep. Proc. IX European Intake Workshop, IGER, North Wyke, Devon, UK., pp. 53-57. Moshtagi Nia, S. A. and K. M. Wittenberg. 2002. Evaluation of n-alkanes as markers for estimation of dry matter intake and digestibility in steers consuming allforage or concentrate-forage diets. Can. J. Anim. Sci. 82:419-425. 43 Newman, J. A., W. A. Thompson, P. D. Penning, and R. W. Mayes. 1995. Least-squares estimation of diet composition from n-alkanes in herbage and faeces using matrix mathematics. Aust. J. Agric. Res. 46:793-805. Oliván, M., L. M. M. Ferreira, R. Celaya, and K. Osoro. 2007. Accuracy of the n-alkane technique for intake estimates in beef cattle using different sampling procedures and feeding levels. Livest. Sci. 106:28-40. Oliveira, D. E., M. Queiroz Manella, L. O. Tedeschi, S. C. da Silva, and D. P. D. Lanna. 2008. n-Alkanes to estimate voluntary forage intake by cattle using controlled release capsules. Sci. Agr. (Piracicaba, Brazil) 65:230-238. Ordakowski, A. L., D. A. Kronfeld, J. L. Holland, B. J. Hargreaves, L. S. Gay, P. A. Harris, H. Dove, and D. Sklan. 2001. Alkanes as internal markers to estimate digestibility of hay or hay plus concentrate diets in horses. J. Anim. Sci. 79:1516-1522. Parker, W. J., S. T. Morris, D. J. Garrick, G. L. Vincent, and S. N. McCutcheon. 1990. Intraruminal chromium controlled release capsules for measuring herbage intake in ruminants – a review. Proc. N.Z. Soc. Anim. Prod. 50:437-442. Piasentier, E., S. Bovolenta, and F. Malossini. 2000. The nalkane concentrations in buds and leaves of browsed broadleaf trees. J. Agric. Sci. (Camb) 135:311-320. Raats, J. G., L. Webber, N. M. Tainton, and D. Pepe. 1996. An evaluation of the equipment for the oesophageal fistula valve technique. Small Rum. Res. 21:213-216. Salt, C. A., R. W. Mayes, P. M. Colgrove, and C. S. Lamb, C.S. 1994. The effects of season and diet composition on the radiocaesium intake by sheep grazing a heather moorland. J. Appl. Ecol. 31:125-136. Sibbald, A. M., G. c. Davidson, and R. W. Mayes. 2000. Effect of dosing regime on intake estimation using the nalkane technique in sheep fed pelleted grass meal. J. Sci. Food Agric. 80:1206-1210. Smith, D. G., R. W. Mayes, and J. G. Raats. 2001. Effect of species, plant part, and season of harvest on n-alkane concentrations in the cuticular wax of common rangeland grasses from southern Africa. Aust. J. Agric. Res. 52:875-882. Smith, D. G., R. W. Mayes, T. Hollands, D. Cuddeford, H. H. Yule, C. M. Malo Ladrero, and E. Gillen. 2007. Validating the alkane pair technique to estimate dry matter intake in equids. J. Agr. Sci. (Camb) 145:273281. Stevens, D. M., J. B. J. van Ryssen, and J. P. Marais. 2002. The use of n-alkane markers to estimate the intake and apparent digestibility of ryegrass and Kikuyu by horses. S. Afr. J. Anim. Sci. 32:50-56. Swain, D. L., M. A. Friend, R. W. Mayes, L. A. Wilson, and M. R. Hutchings. 2008. Combining and active transponder system with sprayed n-alkanes to quantify investigative and ingestive grazing behaviour of dairy cattle in pastures treated with slurry. Appl. Anim. Behav. Sci. 109:211-222. Tilley, J. M. A. and R. A. Terry. 1963. A two-stage technique for the in vitro digestion of forage crops. J. Br. Grass. Soc. 18:104-111. Unal, Y. and P. C. Galsworthy. 1999. Estimation of intake and digestibility of forage-based diets in group-fed dairy cows using alkanes as markers. J. Agric. Sci. (Camb) 133:419-425. Valiente, O. L., P. Delgado, A. de Vega, and J. A. Guada. 2003. Validation of the n-alkane technique to estimate intake, digestibility, and diet composition in sheep consuming mixed grain:roughage diets. Aust. J. Agric. Res. 54:693-702. 44 45 13 18 15 2 13 9 4 Calluna vulgaris Fagus sylvatica Lotus pedunculatus Picea sitchensis Pinus sylvestris Trifolium repens Trifolium subterraneum 0 24 10 0 27 32 Cynodon dactylon Festuca arundinacea Lolium perenne Pennisetum clandestinum Phalaris aquatica Setaria anceps 82 17 7 33 42 11 89 15 16 35 28 3 151 361 63 6 C27 62 21 12 77 129 30 180 116 250 108 21 8 212 13 160 456 C29 74 16 79 103 216 66 243 60 74 124 1 4 55 3 636 136 C31 25 7 195 84 59 91 137 34 10 15 10 2 37 3 458 5 C33 Alkane carbon-chain length 5 2 204 11 2 58 32 4 - 4 4 0 1 0 14 0 C35 28 19 30 104 27 31 64 13 193 18 81 85 28 191 363 23 1-C24 20 3726 9 2628 639 25 56 85 408 143 58 18 2463 89 167 142 1-C26 - 23 - 10 10 - - 17 0 0 1853 2065 0 7 14 196 10-C29 - 45 - 446 101 - 187 4052 281 61 13 14 1327 3 260 25 1-C28 LCOH carbon-chain length Adapted from data in Bugalho et al. (2004), Ali et al. (2005a), and Dove and Mayes (2005). 12 Chloris gayana 1 6 Bromus carthaticus Monocotyledons 2 C25 Brassica oleracea Dicotyledons Species 24 472 0 627 58 0 149 84 4709 1297 114 20 1285 39 298 29 1-C30 Table 1. Concentrations (mg/kg DM) of the major (odd-chain) n-alkanes and the major long-chain alcohols (LCOH; even-chain except for 10-C29-OL) in a selection of dicotyledons and monocotyledons1 Table 2. Concentrations (mg/kg DM) of the major (even-chain) long-chain fatty acids (LCFA) in a selection of dicotyledons and monocotyledons1 LCFA carbon-chain length C20 C22 C24 C26 C28 C30 C32 Acer pseudoplanatus 49 2 293 203 421 278 23 Brassica oleracea 36 29 41 13 16 20 3 Dicotyledons 2 Calluna vulgaris NR 700 556 516 572 445 449 Fagus sylvatica 94 336 111 71 292 16 7 Picea sitchensis 502 657 430 60 48 110 31 Pinus sylvestris 570 1069 159 21 32 63 36 Trifolium repens NR 613 716 608 792 139 41 Heliotropium sp. 267 483 275 151 99 63 36 Stylosanthes guianensis 232 357 540 363 352 398 398 Lolium perenne 1 99 190 174 156 159 88 34 Lolium perenne 2 NR 252 257 187 204 171 74 Cenchrus ciliaris 236 287 331 279 249 235 285 Chloris gayana 241 258 312 223 146 225 521 Hordeum vulgare 283 466 461 270 369 299 170 Monocotyledons 1 Re-tabulated from the data in Ali et al. (2005a, 2005b) and Ferreira et al. (2009a). NR = not reported. 2 46 Table 3. Comparison of measured herbage intakes of herbivores with those estimated using dosed and herbage alkanes Known intake Source Animals/conditions Mean discrepancy ± SE1 Sheep Mayes et al. (1986) Dove and Oliván (1998) Dove et al. (2002) Sibbald et al. (2000) Lewis et al. (2003) Valiente et al. (2003) 36 kg lambs/fresh herbage (g of DM/d) 579 0 ± 4.7 (0%) 30 kg sheep/chaff+sunflower meal 720 -6 ± 6.4 (0.8%) Adult sheep/pre-frozen herbage Adult sheep (Group 1)/pelletted grass meal Lambs (0.3, 0.45 of mature weight)/ forage diets 57 kg ewes/straw+barley grain 914 0 ± 15.5 (0%) 2040 -39 ± 27.4 (1.9%) 2016 -49 ± 28.3 (2.4%) 748 -46 ± 24.0 (6.1%) Goats Giraldez et al. (2004) (g of DM/d) 43 kg goats/fresh herbage, pulse dose of alkane Cattle 808 18 (-2.1%) (kg of DM/d) 8.1 7.9 22.5 -0.1 ± 0.04 (-1.4%) -0.2 ± 0.05 (-1.0%) 0 Unal and Galsworthy (1999) Holstein-Friesian cows (± lactating)/hay+rolled barley Berry et al. (2000) 611 kg Brown Swiss cows/forage:concentrate mix 12.7 0.03 (0.2%) Mann and Stewart (2003) Yearling bulls/fresh kikuyu 6.3 0.1 (1.1%) 457-635 kg cows/meadow hay 6.7 -0.3 ± 0.17 (-3.9%) 233 kg Tuli steers/poor-quality hay 3.2 -0.2 (-6.3%) 380 kg Nelore steers/TMR 5.0 0.1 (2.0%) 490 kg geldings/ryegrass or kikuyu 153 kg donkey/poor-quality hay 149-223 kg donkey/hay+barley straw 228-496 kg horse/haylage (kg of DM/d) 7.2 2.3 3.0 5.4 0.3 (4.2%) -0.1 (-2.7%) 0.4 (13.0%) 0.5 (9.1%) Ferreira et al. (2004) Smith et al. (2007) Oliveira et al. (2008) Equids Stevens et al. (2002) Smith et al. (2007) Cervids (kg of DM/day) 295 kg pregnant wapiti hinds/alfalfa Gedir and Hudson (2000) 9.2 -0.2 (-2.2%) cubes 1 Discrepancy = (known intake – estimated intake), with SE of mean discrepancy where possible. Values are g of DM/d for sheep and goats, kg of DM/d for other species. Values in parenthesis are the percentage over- or under-estimates of known intake. 47 Table 4. Regression1 relationships between known diet composition (X) and that estimated from the alkane profiles of feces and of diet components (Y) for a range of livestock species Species Reference Slope Intercept R2 Significance Cattle Sheep Goats Horses Ferreira et al. (2007a) 0.959 0.014 0.998 NSD Y = X Ferreira et al. (2009b) 0.972 0.006 0.996 NSD Y = X Lewis et al. (2003) 0.990 0.023 0.950 NSD Y = X Valiente et al. (2003) 0.995 0.236 0.990 NSD Y = X2 1.006 -0.025 0.990 P < 0.053 Elwert et al. (2004)* 1.003 -0.002 0.946 NSD Y = X Elwert and Dove (2005)* 0.995 0.002 0.998 NSD Y=X Fraser et al. (2006) 0.907 0.009 0.978 NSD Y = X4 Charmley and Dove (2007) 1.001 -1.986 0.812 NSD Y = X Ferreira et al. (2009b) 0.891 0.020 0.887 NSD Y = X Brosh et al. (2003) 0.998 0.000 0.985 NSD Y = X Ferreira et al. (2009a) 0.983 0.006 0.999 P < 0.055 Ferreira et al. (2009b) 0.857 0.025 0.918 NSD Y = X Ferreira et al. (2007a) 0.925 0.025 0.987 Slope P < 0.055 Ferreira et al. (2009b) 0.923 0.013 0.964 NSD Y = X 1 Regressions shown are either as published in the references shown (*) or are calculated from data therein; for multi-component diets, all components were included in the regression. Unless otherwise stated, fecal recovery corrections were based on treatment mean values. 2 Values for proportion of grain in diet. 3 Recalculated values for proportion of alfalfa in diet. 4 Values are for best solution (M1), without fecal correction because recoveries were complete. 5 Intercept not significantly different from zero and regression constrained through origin not significantly different from Y = X. 48 Table 5. Comparison of actual organic matter (OM) intakes (g/d) by sheep with those estimated using the C32/C33 or C36/C33 alkane pairs following dosing with intra-ruminal alkane CRD, or those estimated from diet composition and the known intake of an alkane/LCOH-labeled supplement Intake method Daily OM intake Difference cf. actual (%) Actual intake 674.1 - C32/C33 702.9 +4.3 C36/C33 699.2 +3.7 Alkane/individual1 679.4 +0.8 Alkane/treatment mean 670.8 -0.5 Alkane/grand mean 683.9 +1.5 (LCOH+alkane)/individual 675.0 +0.1 (LCOH+alkane)/treatment mean 698.8 +3.7 (LCOH+alkane)/grand mean 671.3 -0.4 From alkane CRD From labeled supplement 1 Markers used to estimate diet composition/data used in fecal recovery correction. Values recalculated from Charmley and Dove (2007) and Dove and Charmley (2008). 49 Resultant error in intake estimate (%) 30 20 10 0 -10 -20 -5 -4 -3 -2 -1 0 1 2 3 4 5 Error (%) in fecal output or digestibility estimate; difference (%) in alkane recovery Figure 1. Comparative effect on the error of estimated intake of errors of ± 5% in the estimate of fecal output (■), ± 5% in the estimate of a digestibility of 75% (□), or differences in fecal recovery of ± 5 percentage units in fecal recovery of the alkanes used to estimate intake (Δ). 60 Y = 1.28X - 0.83 Error in intake estimate (%) r2 = 0.986, P<0.001 40 20 0 -40 -20 0 20 40 60 -20 -40 Difference in recovery of alkane pair (% ) Figure 2. Effect of the difference in recovery between the 2 alkanes used to estimate intake on the resultant error in estimated intake. The solid line is the fitted regression, which does not differ significantly from the theoretical relationship (dotted line). Recalculated from Charmley and Dove (2007) and Oliván et al. (2007). 50 1.2 (a) Alkane recovery in ruminant species recalculated from the following: (a) Fecal recovery 1 Sheep: Dove and Oliván (1998; ■); Elwert et al. (2004; □); Elwert and Dove (2005; ▲); and Ferreira et al. (2009b; Δ). 0.8 Cattle: Smith et al. (2007; ●) and Ferreira et al. (2009b; ○). 0.6 Goats: Brosh et al. (2003; ♦ odd-chain alkanes) and Ferreira et al. (2009b; ◊). 0.4 0.2 0 24 25 26 27 28 29 30 31 32 33 34 35 36 1.2 (b) Alkane recovery in equines recalculated from the following: (b) Fecal recovery 1 Horses: Ordakowski et al. (2001; ■ means across diets); Stevens et al. (2002; □); and Ferreira et al. (2009b; Δ) 0.8 Donkeys: Smith et al. (2007; ▲) 0.6 0.4 0.2 0 24 25 26 27 28 29 30 31 32 33 34 35 36 1.2 (c) Alkene recovery data (sheep only) adapted from Elwert et al. (2004; ■), Elwert and Dove (2005; □), and Dove and Oliván (2006; ▲). (c) Fecal recovery 1 0.8 0.6 0.4 0.2 0 24 25 26 27 28 29 30 31 32 33 34 35 36 Alkane or alkene carbon-chain length Figure 3. Effect of carbon-chain length on fecal recovery of alkanes and alkenes of equivalent chain length. 51 1 (a) Faecal recovery 0.8 0.6 0.4 0.2 0 20 22 24 26 28 30 32 LCOH carbon-chain length 1 (b) Fecal recovery 0.8 0.6 0.4 0.2 0 20 22 24 26 28 30 32 34 LCFA carbon-chain length Figure 4. Influence of carbon-chain length on fecal recovery of LCOH (a) recalculated from the data of Ali et al., (2004; ■) and Dove and Charmley (2008; □) and LCFA (b) recalculated from the data of Ali et al. (2004; ■) and Ferreira et al. (2009a; □). 52 500 Estimated supplement intake (g DM/d) y = 1.0066x + 9.0859 R2 = 0.9905 400 300 200 100 0 0 100 200 300 400 500 Known supplement intake (g DM/d) Figure 5. Relationship between known supplement intakes and those estimated using beeswax as an alkane marker, and natural and dosed alkanes to estimate total intake (adapted from Dove and Oliván, 1998). The fitted regression shown did not differ from Y = X. Supplement intakes estimated using alkenes (unsaturated hydrocarbons; Dove and Oliván, 2006) were related to known intakes by the following expression, which also did not differ significantly from Y = X: [Y = 1.057 * X 0.010 (r2 = 0.995)]. 53 Estimated intake (g DM/day) 750 (a) 650 550 450 350 Estimated intake (g OM/day) 350 450 550 650 750 150 300 450 600 600 (b) 450 300 150 0 0 Known roughage intake (g OM/day) Figure 6. Comparison of known forage intakes and the intake of (a) a single forage (Elwert and Dove, 2005) or (b) multiple forages (Charmley and Dove, 2007) calculated using the estimated proportions of forage and supplement in the diet, together with known supplement intake. In neither case, the fitted regression did not differ from the line of equality (solid lines shown). 54 KEYNOTE: ASSESSMENT OF INTAKE AND DIET COMPOSITION OF GRAZING LIVESTOCK H. Dove Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 GEOSPATIAL METHODS AND DATA ANALYSIS FOR ASSESSING DISTRIBUTION OF GRAZING LIVESTOCK D. M. Anderson1 2 USDA-ARS Jornada Experimental Range, Las Cruces, NM 88003 Key words: foraging, geographic information system, global positioning system, landscape utilization, range animal ecology ABSTRACT: Free-ranging livestock research must begin with a well conceived problem statement and employ appropriate data acquisition tools and analytical techniques to accomplish the research objective. These requirements are especially critical in addressing animal distribution. Tools and statistics used to describe the plant-livestock interface must consider that foraging is highly place dependent, foraging is not a random process and livestock foraging mainly occurs in groups. Since animal movement implies consecutive locations, each the result of a host of interactions between the animal and its surroundings, classical agronomic statistics may not provide the optimum analytical tools to address spatially related phenomena. By far the greatest amount of research on quantitative tools to understand the use of space by animals has come from the wildlife community and more recently from ecologists exploring the distribution of both animal and plant species over large geographic areas. However, interest in geospatial data analysis methods is also found among animal-range scientists, primarily because new electronic technologies make it possible to accumulate data from free-ranging animals in seconds or less. Therefore, the rate at which data can now be collected produces inordinate amounts of information that must be correctly analyzed. Consecutive animal locations on a landscape are not statistically independent; therefore, autocorrelation is no longer a theoretical but real concern and must be addressed in animal distribution analyses. This paper presents an overview of research that affects landscape use together with possible geospatial tools that can be used when evaluating free-ranging animal distributions. Geospatial methodology is not trivial nor is it a mature field, but rather is evolving; therefore, the range-animal scientist should work as a team member with other disciplines including, but not limited, to statisticians, modelers, computer scientists, cognitive scientists, physicists, sociologists, and possibly geographers when attempting to understand animal distribution. Literature referenced in this paper is comprehensive in breadth but not exhaustive in detail and thus serves as a beginning point for those wanting to know what has already been researched concerning livestock distribution on landscapes. INTRODUCTION Patterns of behavior that result in the distribution of free-ranging animals over extensive landscapes are poorly understood (Vavra and Ganskop, 1998). Vavra and Ganskop (1998) stated that the greatest frustration in the study of foraging behavior has been the lack of a unifying theory capable of explaining the selective process of foragers and their distribution. Sumpter (2006) suggested that because there are many underlying principles involving collective animal behavior it is unlikely “a single formalism” can describe them all. Distribution is a complex issue to address biologically and mathematically because an animal‟s use of space (its distribution) is related to both internal (biological rhythms and drives) as well as external (environmental) factors (Calenge et al., 2009). Understanding animal production is not in itself adequate to comprehend the complexities of foraging (Morley, 1981). Furthermore, since free-ranging animals are not static either spatially or temporally, the complexity of understanding distribution is no trivial matter. It is unlikely free-ranging animals will use the whole of the space available to them on their own unless stocking is intensive (Arnold and Dudzinski, 1978). Arnold and Dudzinski (1978) further suggested flocks and herds that split into more than a single group may disperse further over the landscape, although research addressing this concept is lacking. Under the direction of a herder or range rider (Skovlin, 1957) livestock are directed to the most palatable forage (Turner et al., 2005). Herding shows great promise as a means to improve animal distribution (Bailey, 2004) and with new innovations such as virtual fencing, “electronic herding” may become practical in the 21 st century (Anderson, 2007). Herding tends to provide foraging animals with a higher quality diet than if animals are left to their own behaviors. Hunt et al. (2007) suggests that management techniques that rely on forcing cattle to use certain areas, such as fencing or developing appropriately located points for drinking water may be more effective than passive approaches that allow cattle considerable freedom to choose where in a paddock they graze. Neither feral (Lazo, 1994) nor domestic cattle (Schmidt, 1969) use paddocks uniformly. Casual observation reveals both under- and overuse areas occur within the same paddock (Coffey, 2009), indicating neither 1 Corresponding author: [email protected]. Mention of a trade name does not constitute a guarantee, endorsement, or warranty of the product by the USDA-ARS over other products not mentioned. 2 57 distribution, an approach normally outside the training of most animal ecologists. In designing foraging experiments, it is necessary to consider biological, practical, and statistical implications (Bransby, 1989). Furthermore, software development is never trivial (de Smith et al., 2007). Therefore, it should be immediately obvious that geospatial data analysis of groups must be considered from the standpoint of the biologist, cognitive scientist, psychologist, physicist, mathematician, computer scientist, as well as the geographer and sociologist (Schellinck, 2008; Calenge et al., 2009). Assembling teams that communicate and work together effectively to address free-ranging animal behavior is a challenge. Just as early work on animal behavior was not conducted by scientists, trained as ethologists (Squires, 1975), many biologists of today are attempting to model and understand the plant-animal interface without a complete mathematical understanding of the models. Conversely, some modelers lack a complete biological understanding of what they are attempting to model (Loewer, 1989). There are other reasons plant-animal interface research is challenging. First, no standardized protocol in experimental grazing studies has been advocated (Amer. Soc. Agron. et al., 1952) making comparisons among studies challenging. Second, the importance of proper grazing management is often ignored by development and conservation communities (Turner et al., 2005). Third, although livestock husbandry is one of the oldest industries, little from the fields of monitoring and processing used by other industries has been adopted because the factors required for biological monitoring are inherently variable and unpredictable (Frost et al., 1997). Forth, often if freeranging animal behavior is considered important, it is assumed the individuals that make up a herd differ little in behavioral characteristics and hence will respond similarly to management practices and actions (Hunt et al., 2007). Fifth, although animal behavior is now recognized as pivotal in the foraging process and the use of space (Hancock, 1953; Stricklin and Kautz-Scanavy, 1984; Senft et al., 1987; Coughenour, 1991; Stuth, 1991; Bailey et al., 1996), history bears out this has not always been the case. Animal behavior has only been recognized as a full-fledged part of animal science since the mid-1950‟s (Curtis and Houpt, 1983). Early foraging research either overlooked or chose not to address the complexities associated with animal behavior. Burns et al. (1989) recommended that the following 4 measurements be made in all grazing experiments: 1) estimate herbage mass; 2) estimate green leaf; 3) estimate diet quality; and 4) determine herbage density, but nothing was mentioned concerning animal behavior. Sixth, guidelines on how to conduct grazing studies frequently do not address how to document animal behavior in foraging studies (Henderson, 1959; Davies, 1961; Morley, 1981; Cook and Stubbendieck, 1986). wild animals (Seaton, 1909) nor livestock use landscapes uniformly. This fact has been substantiated by monitoring consecutive 24-h periods of free-ranging cows (Anderson, 2007) as well as by investigating fecal deposition patterns (Tate et al., 2003). One of the largest risks in attempting to address large scale phenomena such as animal dispersion is that it may not be possible to identify processes in smallscale studies that are crucial on a larger scale (Marsh and Jones, 1988). However, Shiyomi and Tsuiki (1999) were able to find free-ranging cattle in small herds to be aggregated and not randomly distributed. This paper presents selected literature of past research directed at understanding the multifaceted phenomenon called free-ranging animal distribution. Because of the fast moving changes in the software industry and the continued development of new tools and analytical techniques this manuscript should only serve to focus further research on what questions have yet to be answered and aim future research towards increasing our understanding of freeranging livestock distribution and how to manipulate livestock distribution to foster ecologically based management. BACKGROUND Foraging is a complex process (Reid, 1951) and remains a major issue facing rangeland managers (Bailey et al., 1996). Animals distribute themselves based on multi-level responses to their perceived environment; however, the effect of social interactions on spatial choices has largely been ignored. Although most livestock species are gregarious, it is not known if spatial decisions are made collectively or if only 1 or more animals make these decisions (Bailey and Rittenhouse, 1989; Bailey, 1995). Huber et al. (2008) observed that purebred animals (2 groups with 50 Ankole cattle) had closer nearest-neighbor distances and engaged in more non-agonistic social interactions than crossbred cows (2 similar groups of 50 Holstein x Ankole cattle). An animal‟s preference for heterogeneously distributed resources results in the heterogeneous usage of space, which is a joint function of preference and accessibility (Matthiopoulos, 2003). Although wildlife sciences have led in the development of tools and analytical techniques to analyze free-ranging animals, we remain relatively ignorant about an animal‟s use of space even with the many technological advances in the field of radiotelemetry (Marzluff et al., 2001). Horne et al. (2008) suggested foraging is complex because the use of space by animals is a multivariate process and multivariate statistics require a solid understanding of matrix algebra. Few grazing researchers are fully trained in experimental design and data analysis (Hart and Hoveland, 1989). Moorcroft and Lewis (2006) noted that animal ecologists have often used purely descriptive statistics to describe animal movements rather than using models that may require a more rigorous understanding of mathematics, including differential equations. For example, Paraan and Esquerra (2006) have derived exact expressions involving the Poisson distribution for the first 4 moments of displacement within a THE IMPORTANCE OF BEHAVIOR (EXAMPLES) Most US grazing experiments have used beef steers (Burns and Standaert, 1985); yet, the results from such studies are often applied directly to the management of 58 female animals. Anderson and Holechek (1983) found heifer diets had 4% greater forb content than steer diets, which was the likely explanation for heifers having a 3% greater (P < 0.05) dietary CP content than steers (14%). If nutritional data from 1 class of animal should not be extrapolated directly to another animal class when making nutritional management decisions, then same should be true for applying behavioral data to management. Socialization among animals also has been largely ignored, yet this phenomenon has been shown to affect animal distribution. Bond et al. (1967) suggested that interspecific associations could produce an important variable in grazing studies. The observation that spatial distribution of mixed species groups can be a positive in management prompted an entirely new research direction involving mixed-species stocking and flerds (Anderson et al., 1988). Another often overlooked variable is animal numbers. A simplistic example will serve to illustrate the point. If 20 cows are stocked on an area for 2 days, it equates to 40 animal-unit-days (AUD) of use, an animal unit consisting of a cow-calf pair with the cow weighing approximately 454 kg (Havstad et al., 2004). However, we could arrive at the exact same number of AUD if the same area were stocked with 2 cows for 20 days, although the set of biotic and abiotic factors affecting the plants, animals and their subsequent use of the area (distribution) would be entirely different. The “group” behavior of 20 cows is intuitively different than that of 2 cows, but there is currently a dearth of research on the effect of group size and its affect on distribution. Replication remains a concern and a point frustration when analyzing plant-animal interface studies (Bonham, 1986). Replication is essential and should be done where appropriate to elucidate treatment differences. However, similar to landscape ecology in which there is a conflict between the need to replicate and the need to study processes (Hargrove and Pickering, 1992), some behavior studies seem to defy classical replication. Anderson et al. (1987) designed a study involving socialization of sheep to heifers in pen confinement. Those authors found that 1 particular randomly assigned heifer was physically abusive to the sheep, and as a result, the sheep did not become socialized to the heifers. Despite repeated attempts made to replicate the trial, neither the intra- nor interspecies responses documented in the initial “group” were ever again observed. This outcome suggests that simply randomly assigning a mix of animals may produce a group dynamic that thwarts socialization of the group, and highlights the complexity involved in attempting to conduct free-ranging animal behavior research. Although all the classic factors (age, gender, past handling history, and numbers) were controlled by Anderson et al. (1987), there remained a “psychological aspect” that was not possible to replicate. The take home lesson was one must be concerned with much more than simply numbers to understand animal groups. Mirabet et al. (2007) developed an animal grouping model for fish that accounted for numbers of individuals as well as those influential neighbors that impact the grouping, which too should be considered when attempting to model cattle groupings. WHAT CONSTITUTES A GROUP? An ethology dictionary by Immelmann and Beer (1989) did not provide a numerical definition for “group.” Gerard et al. (2002) suggested group fusion and fission is dynamic with group size and is often suboptimal. Furthermore, cooperative and non-cooperative behaviors apparently change with group size (Chase, 1980) and may be related to predator avoidance strategies (Fortin et al., 2009). Haydon et al. (2008) found in a herd of 100 elk studied over 4 years that the spatial and social context an individual experiences influences the amount of time spent in different movement states, suggesting that social influences impact spatial distribution. Spacing and social behavior also have been investigated using animats (Stricklin et al., 1995). Many behavior models that address animal groupings consider predation to play a large part in this phenomenon (Reluga and Viscido, 2005). Although this has a theoretical and historical interest, the selection pressures of domestication most likely have reduced the importance of this behavioral trait in explaining group formation in domestic livestock. Today, livestock groupings are almost totally under man‟s control, which alters natural tendencies. The motion of individuals in a group is assumed to be the combined result of both density-independent as well as density-dependent decisions (Gueron et al., 1996) and may even be related to physiological variables, such as the size of the neocortex as in primates (Dunbar, 1992). Furthermore, the resource dispersion hypothesis (RDH) asserts that group living may be less costly where resources are heterogeneous in space or time (Johnson et al., 2002). Thus, the aggregation of individuals into groups is a complex phenomenon and rather than providing an explicit definition for animal groups we frequently rely on our familiarity with common animal aggregations (i.e., a flock of sheep or a herd of cows) instead of addressing the mechanisms specific species use to aggregate (Schellinck, 2008). The first theoretical description of aggregate behavior may have been given by Parr (1927) from observation of chub mackerel (Scombrus coligs). Schellinck (2008) suggested the movement of groups involves the use of an animal‟s sensory and perceptual abilities to successfully aggregate because there are no other apparent aggregation mechanisms available to animals; thus, movement follows a loop of information beginning with perception that leads to cognition followed by action and then returning back to perception. Warburton and Lazarus (1991) used computer simulation to evaluate attraction-repulsion distance functions to evaluate inter-individual social cohesion in animal groups. An optimal animal size group for livestock that remains cohesive under all situations has yet to be determined (Stricklin and Mench, 1987) and probably does not exist, yet cattle herds are composed of social subgroups that remain somewhat stable (Sowell et al., 2000). Bransby (1989) suggested paddocks be stocked with at least 3 to 5 animals throughout a grazing period and that a completely random design with rectangular paddocks oriented in the direction of known or expected variation be used to minimize experimental error. He also noted that an experiment with less than 8 paddocks and 4 error degrees of 59 including electrocardiogram (ECG), electroencephalogram (EEG), and respiratory signals from unrestrained animals is now possible (Lowe et al., 2007). Although video techniques have also been used with livestock (Sherwin, 1994), video is not the tool of choice when attempting to document foraging studies. However, video techniques may be useful at locations where an entire group may congregate, such as at drinking water locations or where mineral is provided. It is worth noting that a camera should be placed on a tower for two reasons. First, the electronic equipment is removed from the direct path of animals that could damage the equipment. Second, it positions the equipment at a height that increases the field of view and the amount of information from the surroundings that may be useful in interpreting the data of individuals. If video is used in conjunction with other electronic data acquisition techniques, a better understanding of group behavior may result, especially if not all the animals within the group are monitored or instrumented with individual devices. The attachment of devices to monitor distance traveled may have originated with pedometers (Anderson and Urquhart, 1986). However, the question of “how is the observer influencing the observation” must always be asked when devices are attached to animals. In some cases, there may be a negative effect on behavior because the mass of the device may alter an animal‟s foraging range (Wilson et al., 1986). Technology in the form of telemetry removes observer influence, but there remains a dearth of research addressing how present-day instrumentation affects “normal behaviors” (Ropert-Coudert and Wilson, 2005). Collars that self-adjust to changes in growth of an animal have been designed for wildlife species, including pronghorn fawns (Keister et al., 1988). Generally, behavior altercation is normally not observed with domestic livestock because of the small mass of the device relative to the animal‟s mass and method of attachment (usually collars or neck-body saddles) does not seem to have negative long-term effects on behavior. Animals soon (few hours to a few days) accept attached devices without producing any unusual visual behavior. With the advent of microelectronics began the focused study of foraging strategies among free-ranging animals (Gordon, 1995). The attachment of sensory and recording systems to animals has been termed “bio-logging” (Naito, 2004). Modern devices record animal data and are capable of recording data about the external environment through which the animal is moving (Boehlert et al., 2001). The smallest bio-loggers available presently have been developed for avian research (Ropert-Coudert and Wilson, 2005). von Hünerbein and Rüter (2001) described a GPS flight recorder for homing pigeons weighing less than 50 g. Most livestock researchers focused on animal distribution use GPS technology. This navigation and precise-positioning tool was developed by the United States Department of Defense in 1973 for locating objects (Herring, 1996). Other techniques, including the use of very high frequency (VHF) collars or tags for locating animals predates GPS (Le Munyan et al., 1959; Rempel et al., 1995). It should be mentioned that many of the tools and techniques used today with free-ranging domestic animals had their roots in the wildlife sciences. To learn more about freedom (df) is hard to justify regardless of design. On California rangelands, Harris et al. (2007) found the optimal animal number to be 3 to 6 cows. Hernandez et al. (1999) found 98% of feral cattle on the Chihuahuan Desert in Mexico resided in groups of 1 to 20 while domestic cattle were never found in groups of less than 30 and 90% of the time the cows were in groups greater than 50. Lazo (1994) found for feral cattle in Spain, the number of mature cows in each of 4 herds was between 13 and 32 during autumn. With the advent of global positioning system (GPS), researchers must avoid drawing conclusions from data coming from too few animals, not only from a statistical stand point but more importantly, because of missing individual behaviors that ultimately affect group behaviors (Hunt et al., 2007). Girard et al. (2006) found that for determining moose habitat, it was more important to sample more animals than to increase the number of locations sampled for each animal. Because of the inherent variability among cattle, 5 animals per treatment are probably not adequate when investigating beef cattle behaviors (BishopHurley et al., 2007). Between 10 and 20 animals per treatment seems more realistic, but this will depend upon the behaviors being investigated. If possible, a preliminary trial should be conducted to investigate sample size prior to conducting a larger study. Group size affects social strategies in farm animals (Estevez et al., 2007) and cognitive mechanisms in both wild and domesticated animal species (Croney and Newberry, 2007). Unfortunately, most studies on group size have focused on penned (Christman and Leone, 2007) rather than freeranging animal groupings. A question yet to be answered is how does manipulating free-ranging group size while maintaining a constant stocking density affect behavior? EQUIPMENT One of the first decisions a researcher must make after the objective of the research project has been established is the equipment to be used to collect the data. There are a number of sensing systems available, each suited to answer specific questions under particular circumstances (Table 1). Foraging animals may be observed directly or measured with mechanical or electronic devices (Hart and Hoveland, 1989). The earliest information on time spent foraging was obtained through visual observation over a period of days (Gary et al., 1970). Regardless of the equipment used to locate animals, including the human eye, there will be error in determining the animal‟s exact location (Springer, 1979). Vibracorders, devices developed to monitor the operating times of logging trucks, may have been the first commercially available device adapted to free-ranging livestock providing a time stamped signature of the temporal aspects of foraging (Allden, 1962; Stobbs, 1970). Canaway et al. (1955) developed a recorder that nonsubjectively recorded walking, lying, standing, grazing and cudding. Young (1966) developed equipment for recording jaw movement patterns, and O‟Shea (1969) developed an animal mounted device for recording jaw movements and elapsed time. Devices have also been built for monitoring physiological parameters for use in extensive animal monitoring (Puers et al., 1993). Transmission of data, 60 equipment was bulky and required manual processing. By combining electronic data acquisition using systems (i.e., GPS) with direct observation, daily behavior itineraries have been accurately determined (Schlecht et al., 2006). Using GPS fixes between 1 Hz (Anderson, 2007) and 4 Hz (Swain et al., 2008) to evaluate foraging locations appears promising. The optimal time interval between recordings remains an issue in the geostatistical analysis of GPS trajectory data (Hengl et al., 2008). Remple and Rodgers (1997), using differential correction, were able to decrease GPS location error from 80 to 4 m (P < 0.0001) and the range of 25-75th percentile location error from 74.3 to 5.0 m. Location error in radiotelemetry data lowers the power of statistical tests (Findholt et al., 2002). Therefore, Zimmerman and Powell (1995) advocated for researchers to use statistics derived from the linear distances between actual and estimated locations of test transmitters to estimate location error. Precision agriculture (PA) has focused on understanding agronomic spatial variability in soils and crops (Stafford, 2000); however, the sensors used on arable crops are different from those necessary to discriminate plant species on grassland (Schellberg et al., 2008). For example, texture features and spectral characteristics of plant species may be sufficient for accurate and fast identification of forbs, whereas plant shape analysis may be more important on arable landscapes. Determining gap spacing may be crucial in grasslands. Lock et al. (2004) successfully calculated size and shape of artificially created gaps in turf grass. Landscape composition may have a larger impact on foraging (utilization) patterns than grazing management (Senft, 1989). The use of high resolution spectral sensors and photography has a long history (Deering et al., 1975; Flynn et al., 2008). Furthermore, this approach appears to offer economic (Pickup et al., 1994) as well as biotic (Starks, et al., 2004) and abiotic (Jackson, 1986) advantages for landscape monitoring involving land degradation (Graetz, 1987). Patch dynamics appear to be the most recent focus to understanding ecological processes (Ludwig et al., 2000; Butler, 2002), yet their role in foraging herbivores is less well understood because herbivores tend to exploit food resources that are generally continuous (Searle and Shepley, 2008). Swain et al. (2008) suggested that high duration GPS fix rates could be used to derive speed histograms, and if used in conjunction with low fix rate data, may establish the probability of an animal visiting a given patch within a landscape. Introducing technology into the equation for determining animal distribution adds additional concerns even though technology offers many opportunities (Cox, 2002). One of the challenges when using GPS data may be to accurately differentiate between resting and foraging because foraging has a very low forward trajectory due to the inherent location error of current GPS devices (Schlecht et al., 2004). Selecting a high GPS fix rate and carefully selecting data sets with low Position Dilution of Precision (PDOP) values may greatly enhance the accuracy of activity assessment. Graves and Waller (2006) suggested that an analysis of the pattern of missing fixes should always be part of any trajectory analysis because the radiotelemetry and its use in wildlife science, the text by Amlaner and Macdonald (1980) is considered seminal. Very high frequency radio tags were first deployed in wildlife studies during the late 1950‟s (Kenward, 2001a) and satellite tracking of animals began in the early 1970‟s (Kenward, 1987) when location accuracies were in the 5 km range (Gillespie, 2001). One possible reason equipment developed for wildlife was not simultaneously used with domestic livestock during its early history was that early wildlife equipment was adequate for monitoring very “coarse-grained” movements over very large areas, but the technology had limitations for accurately monitoring movements in relatively small areas, i.e., paddocks or finer-scale temporal scales (Keating et al., 1991). Accuracies of 5 m true position are now common under open sky conditions with consumer-grade GPS receivers (Wing et al., 2005), and with more sophisticated receivers, 0.01 m accuracy is possible (Bajaj et al., 2002). Hyperbolic telemetry can improve the accuracy of VHF data to within a few meters (Rodgers, 2001), but differential correction can do the same for GPS data (Hurn, 1993; Rempel and Rodgers, 1997). Moreover, Kenward (2001b) pointed out that GPS receivers remain relatively expensive. A recent effort to develop a small-animal satellite tracking system, named International Cooperation for Animal Research Using Space (ICARUS) will employ radio technology on a near-earth orbit satellite for tracking animals (Wikelski et al., 2007). Using GPS does not come without challenges. Canopy cover can attenuate GPS signal strength when tracking wildlife (Dussault et al., 2000; DeCesare et al., 2005). Antenna position with respect to the sky can have a significant impact on number of GPS fixes that are missed (D‟Eon and Delparte, 2005; Heard et al., 2008). However, Hiroaki and Takaaki (2007) found that in Brazil even if animals wore GPS equipment with their antennas pointed downward the units still received radio wave signals without significant problems. They attributed this to their sampling position being near the equator. Bolstad et al. (2005) found that a forest canopy produced significant differences in the mean positional error of GPS receivers whereas no statistically significant differences were observed under open locations. Furthermore, GPS data coming from different devices may not necessarily be identical, which can lead to challenges when comparing data among studies. Buerkert and Schlecht (2009) found that 3 unnamed commercial GPS devices programmed to capture fixes every 15 sec produced a similar (P > 0.05) number of position fixes when operating in open terrain but when the terrain was obstructed, the 3 devices produced a different (P < 0.001) number of fixes. This was attributed to factory settings of the GPS receivers, which evidently placed specific weights on signal reliability and trigonometric properties. The earliest studies to track wildlife (Cochran et al., 1965) and free-ranging cattle (Petrusevics and Davisson, 1975) required researchers to build their own equipment. Techniques have become more sophisticated and may include the use of stable isotopes to track animal migration (Hobson and Wassenaar, 2008). Canaway et al. (1955) may have been the pioneers in adapting telemetry to foraging animals even though their 61 frequency of data acquisition can affect the path the animal traverses (Schwager et al., 2007). environment. Furthermore, animals respond to environmental heterogeneity at different scales and in different ways (Johnson et al. 1992). Other obvious factors affecting animal movement are nutrients (this includes water), forage quantity and quality, as well as weather (Ehrenreich and Bjugstad, 1966) and other abiotic factors as nebulous as slope and aspect (Turner et al., 2001). These factors are all integrated through the physiological and mental well being of the kind or mix of animals under consideration. Foraging decisions are a balance between motivation to eat and satiation (Baumont et al., 2000) and they usually do not take place in isolation but in groups that vary with season and forage condition (Dudzinski et al, 1982). Furthermore, Wolff (1993) suggested mammal dispersal distances may not be completely explained by resource competition. Over a 3 year study involving 29 Bos indicus cows, certain individuals were found to form preference partners and these preference partners were also observed among calves (Reinhardt and Reinhardt, 1981). In Montana foothill rangeland, cows tended to have preference partners when the herd size was less than 30, but when the herd size was greater than 30 there was no evidence that cows had preference partners (Pollak et al. 2008). As a result of these associations, foraging animals are responsible for creating spatial patterns of sward heterogeneity that change throughout utilization (Rook et al., 2004). Feeding rate may be a valuable tool for assessing the degree to which an animal is constrained by its social environment (Nielsen, 1999). In general, the time spent foraging will lengthen in response to a decline in available herbage (Hodgson, 1933) and animals foraging on a heterogeneous resource will usually select a diet higher in quality than the standing-crop can offer (Baumont et al., 2000). Tribe (1980) stated that the main challenge in future grazing management research will be to improve the efficiency of animal production from pastures. Accompanying this is the need for statistical designs that encompass the animal and its behavior (Manly et al., 1993). Rutter (2007) stated that in order to understand the spatial dynamics of foraging behavior, we need to combine information on animal position, foraging behavior and vegetation maps. ANIMAL DISTRIBUTION If stocking rate is the first variable to get correct when managing free-ranging animals (Walker, 1995), surely obtaining proper forage utilization by managing animal distribution must be the second most important principle of grazing management. What drives animal distribution and how to manage it continue to be major research foci (Bailey et al., 1996; DelCurto et al., 2005; Anderson, 2007). An early review by Brown and Orians (1970) investigated proximate causes, ecological consequences, and adaptive significance of spacing patterns in mobile animals. Foraging distribution is heterogeneous, especially in large paddocks on extensively managed arid, semiarid, and mountainous rangelands (Coughenour, 1991; Bailey et al., 1996). Wildlife biologists have had the ability to track and map free-ranging animal locations and movements for decades even though the early data were relatively crude, inaccurate, and imprecise (Sampson and Delgiudice, 2006). The rich history of tracking wildlife has been documented by Kenward (2001a). More recently, remote sensing has been used for estimating terrestrial animal distribution and diversity (Leyequien et al., 2007). Tracking animals is mainly a research endeavor at the present time (Schellberg et al., 2008). Determining what affects distribution is not a trivial pursuit. Between 1926 and 2009, at least 68 factors were identified that affected animal distribution (Table 2) either alone or in combination and this list is far from complete. Macdonald and Johnson (2001) believed animal dispersal was one of the least understood aspects of animal ecology in terms of conservation biology. Furthermore, much of what is understood about mammalian dispersal comes from the study of small mammals (Stenseth and Lidicker, 1992); however, dispersal behavior among large mammals may not be identical (Sinclair, 1992). Furthermore, many approaches used to address free-ranging livestock dispersal were borrowed from wildlife research, and while principles involved in wildlife dispersal may be applicable to livestock, dispersal within fenced paddocks requires much discernment when making such linkages. Early studies identifying factors affecting animal distribution patterns, especially foraging, were primarily descriptive (Senft et al., 1983) and involved following animals and periodically recording locations on pasture maps (Dean and Rice, 1974). This approach was followed by using light aircraft to determine cattle distributions (Low, 1978; Low et al., 1981). Such tracking usually took place during early mornings, because behavioral studies indicated that cattle were most likely to be in the vegetation community in which they would spend most of their time foraging. It was also assumed animals would take the shortest paths from their grazing areas to water (Pickup and Chewings, 1988). More recently, Fleming and Tracey (2008) indicated that counting as well as recording animal behaviors from aircraft was fraught with errors. Animal movements are the result of complex interactions between an individual animal and its internal (Loeb, 1918) as well as its external (Forester et al., 2007) COMBINING GEOGRAPHIC INFORMATION SYSTEM AND GLOBAL POSITIONING SYSTEM INFORMATION It is impossible to find a single succinct and clear definition of what constitutes a Geographic Information System (GIS). Geographic information system is a technological field that began in the late 1960‟s and early 1970‟s that integrates hardware and software for capturing, managing, analyzing, and displaying geographically referenced (spatial) data that can be referenced to locations on the earth. Wong and Lee (2005) suggested beginners in GIS may wish to consult online training courses such as those sponsored by ESRI (http://www.esri.com/). Those authors suggested their book called Statistical Analysis with ArcView GIS (Lee and Wong, 2001) was the first serious 62 accuracy, yet neither data model is intrinsically superior and persons who have a grasp of both tools are equipped to develop the most efficient and accurate analysis (Price, 2006). In general, the raster approach is better suited for depicting areas with inexact boundaries and can quantify spatial variability within polygons but requires a large amounts of disk storage; this is in contrast to the vector approach that accurately represents points and linear features having exact boundaries but can only apply an attribute uniformly to an entire polygon and requires a large amount of memory to perform spatial analyses of the data stored in this format (Johnson, 1990). Pickup and Chewings (1988), using band 5 Landsat imagery in the visible red part of the spectrum (0.6-0.7 µm), spatially rectified if necessary, were successful in determining the reduction in percentage of vegetative cover as a result of foraging. Processes that regulate foraging strategies are neither linear nor density dependent across all scales. Extrapolation of known plant-animal interactions across scales is questionable; for example, how do measures of animal productivity on small trials relate to production outcomes in large paddocks (Roshier and Nicol, 1998). What works to improve distribution at one scale should not be assumed to improve distribution at all scales; thus, a combination of techniques effective at different scales seems appropriate (Hunt et al., 2007). Pickup and Chewings (1988) pointed out that scale is important when different data are recorded at different scales. They used Landsat pixels giving a resolution of 57 x 79 m for mapping vegetation while their cattle distribution data were only accurate to about 0.25 km; hence, there was little point in modeling at the Landsat pixel scale given the spatial resolution of the cattle data. Had the Landsat data been aggregated into larger pixels by re-sampling and smoothing, a significant loss of information would have occurred that would have adversely affected the accuracy of the vegetation classification system they were using. Karnieli et al. (2008) found that a pixel size of 171 m was small enough to monitor the effects of livestock foraging gradients in concentric circles out from drinking water for 6 km. It is beyond the scope of this paper to discuss the complexity of pixels and their optimum size except to say that geospatial research must be designed such that field observations match what the satellite “sees” (Cracknell, 1998; Weber, 2006). Semivariogram analysis of California chaparral indicated that 6 m pixels would be optimum for retaining most spatial variation of shrubs (Rahman et al. 2003), however no one pixel size fits all PA management requirements. attempt to integrate spatial analytical and statistical functions with GIS. Textbooks on the basic mathematics and statistics relevant to GIS (Bluman, 2003; Allan, 2004; Dale, 2005; Longley et al., 2005; de Smith, 2006; de Smith et al., 2007) as well as many advanced texts (Haining, 2003) are also available. Methods have been proposed to assess error associated with data in presence-absence matrices (Murguia and Villasenor, 2000). Roloff et al. (2009) points out that most people using GIS data did not collect the GIS data and therefore may not be aware of the data‟s specific limitations. Geographic information system datasets or themes include both spatial and attribute data (Wong and Lee, 2005). deSmith et al. (2007) pointed out that using one GIS package on a given dataset can rarely be exactly matched to results produced using any other package or through hand-crafted coding, yet one product is not necessarily better or more complete than another. Combining GPS and GIS can provide a useful tool for quantifying foraging distribution of sheep and goats (Kawamura et al., 2005) as well as cattle (Barbari et al., 2006; Brosh et al, 2010). Together GPS and GIS provide a powerful toolset for studying free-ranging animal distribution. Commercial products rarely provide access to source code or full details of the algorithms employed even though this information would aid greatly in understanding reproducibility and further development (de Smith et al., 2007). In addition to the software packages listed in Table 3, de Smith et al. (2007) list 46 GIS software packages, some of which are free. Trade magazines such as Geocommunity (http://www.geocomm.com/) provide ad hoc reviews of GIS software, especially new releases. Choosing the “best” tool is up to the user based on the specific questions to be addressed, and at present there are no standardized tests for the quality, speed, and accuracy of GIS procedures (de Smith et al., 2007). Johnson (1990) indicated GIS programs store and manipulate data as either rasters (cells or pixels) or as vectors (points, lines or polygons). Vector and raster refer to two different kinds of mapping methods (Wade et al., 2003). Geographic information system products should support at least 2D mapping (display and output) of raster (grid-based, which is widely used in environmental sciences and for remote sensing) and (or) vector (point, line, or polygon) data with a minimum of basic map manipulation facilities (de Smith et al., 2007). When derivative measurements are used to combine raster and vector data, a conversion of all the data to either one or the other is required for processing (Wade et al., 2003). Although locational resolution for raster-based systems can be improved by reducing pixel size data, volume will increase (Shaw and Atkinson, 1990). Ludwig et al. (2004) found that a 30 to 100 m pixel size from Landsat imagery is adequate to assess the condition of rangelands in Australia along foraging gradients extending from drinking water locations. Conversion of vector data into a raster format may create the modifiable areal unit problem (MAUP; Dark and Bram, 2007) that involves scale and zonation effects that can have a definite impact relative to the analysis of spatially explicit data usually present in various types of large-scale spatial data analysis. It has been assumed that raster calculations were faster whereas vector methods provided higher PRACTICES INFLUENCING GEOSPATIAL ANALYSES The location of resources such as shade and a feeding site can have a major impact on the spatial distribution of cattle (Dolev et al., 2008). Although changing mineral location will cause free-ranging cattle to move, Ganskopp (2001) concluded moving water was more effective than moving salt because the impact of the new salt location waned after about 2 days and the cattle began drifting back towards their previous location. Barnes et al. (2008) suggested that distribution can be improved by switching 63 incorrect conclusions regarding whether relationships between variables or factors are real (Wong and Lee, 2005). For example, the positions of individual cows in a herd are not independent observations but are auto-correlated because animals tend to move together, precluding the use of cows as replicates in parametric statistical tests (Harris et al., 2002). However, Bailey (2010, personal communication) suggested that if a herd of ≥ 30 cow‟s break into several small groups and the individuals that compose each of the smaller groups change daily then it may be possible to consider each member in the group as an experimental unit. However, it is more likely that animals form groupings with companions rather than strangers that are stable over time to meet their social needs (Brunson and Burritt, 2009). The positions of sequentially collected animal locations may be serially correlated (Swihart and Slade, 1985) and as such, use of these locations as the sample size would constitute pseudo replication (Hulbert 1984). Cliff and Ord (1981) noted there are simple indices to evaluate the level of spatial autocorrelation in statistical analysis. Graphical methods for presentation of errors and correlations in errors in spatial data are not fully developed, yet a modeling procedure framework for handling variables recorded at different resolutions has been recommended (Buckland and Elston, 1993). Christman and Lewis (2005) highlighted the importance of knowledge of the spatial arrangement of the group before applying a particular statistic because different test statistics may behave quite differently under the same conditions. There are no universal one-size-fits-all rules that apply everywhere and at every scale; therefore, data from different scales or spatial resolution levels do not yield the same statistical results (Wong and Lee, 2005). Morales and Ellner (2002) suggested that the main challenge for scaling up animal movement patterns resides in the complexities of individual behavior rather than in the spatial structure of the landscape. Classical parametric statistics not only describe how values are distributed (descriptive) but are used to describe or compare a population (inferential) (Wong and Lee, 2005). With classical statistics 1) samples are drawn from populations in which the attribute would form a bell-shaped curve, 2) the population variance is homogeneous within the bounds of random variation, and 3) measurements to be analyzed are continuous with equal intervals (Lehner, 1979). Statistical regression fits these assumptions and has probably been the most common tool used in animal science for population data analysis (Jones and Langemeier, 2009). Jones and Langemeier (2009) noted that data acquisition under less tightly controlled experimentation, as with free-ranging animal studies, will most likely lead to severe violations of the underlying assumptions of the classical linear regression model. Senft et al. (1983) were the first to use regression as a means of predicting patterns of cattle behavior, and observed nonlinear mathematical relationships among behavior and pasture characteristics. Their approach lacked both the derivation and model data to incorporate effects of herd structure and social interactions and their impact on spatial behavior, especially large herds. Yet this approach had the important property of boundlessness (i.e., not being site-specific). Subsequent from extensive to intensive stocking; however, this level of intensification requires fencing which is costly to purchase and time consuming to construct. Cattle may not necessarily choose to forage in areas with the greatest quantity of standing crop but rather may prefer areas dominated by annual grasses and forbs with lower biomass (Hunt et al., 2007). James et al. (1999) found that uniform heavy foraging across the whole landscape could be detrimental to some species native to the landscape. Senft et al. (1983) found that cattle used as little as 30% of a 125 ha pasture and approximately 38 ha received over 60% of the foraging. If this concept of less than non-uniform use of an area is universally true, then stocking rate needs to be calculated based on the areas actually utilized and not the entire pasture. A first attempt at refining stocking rate calculations may be to base it on the area of the landscape on which animals actually traveled and did their foraging (Anderson, 2007; Table 1) rather than assume the entire area will be used uniformly. If this approach were considered to refine the accuracy in calculating stocking rate, one may find predictive relationship between the actual area used and the total area available. Drinking water developments have been advocated on arid and semi-arid landscapes to promote more uniform spatial foraging; however, based on the research of Li and Wang (2006) it seemed that animals distribute themselves out from the center of a resource point in concentric circles. Water points tend to be a focus point for cattle in paddocks 9 to 57 km2 (Hunt et al., 2007). Stafford-Smith (1988) suggested water points were seminal in developing foraging models to address cattle and sheep distribution on arid landscapes; however, piosphere (Washington-Allen et al., 2004) models must take into account that utilization of vegetation near water is different for wildlife than for livestock (Thrash and Derry, 1999). DATA ANALYSIS Statistical tools involve techniques for collecting, analyzing, and drawing conclusions from data (Snedecor and Cochran, 1967), and although spatial statistics are strongly based on classical statistics, the data are also spatially referenced (Wong and Lee, 2005). Spatial autocorrelation is essentially the nature of geography and consequently will almost always be present in spatial data, in contrast to data analyzed with classical statistical techniques that assume observations and associated data lack spatial dependency or autocorrelation (Wong and Lee, 2005). There is now general acceptance that individual locations in spatial data lack statistical independence (Kenward, 2001b). Kernoham et al. (2001) emphasized the importance of sample size and autocorrelation of data in estimating home ranges and animal movements. Griffith (1987) offered a detailed analysis of the impact of spatial autocorrelation. The preferred method to quantify autocorrelation is to determine the Moran coefficient (MC), and although spatial autocorrelation is well understood with linear models, it is not well understood with generalized linear mixed models (Griffith, 2009). Overall, autocorrelation does not affect the accuracy of estimating parameters in inferential statistics; however, it may cause 64 Miller 1965), it most likely is more complex. Fractal geometry, although not widely used to characterize and model animal movement paths, has been used to analyze spatial location data using Lévy flight modeling and generalized entropy (Hagen et al., 2001). Fractal analyses may only work well in situations where the fractal dimension of a movement path is scale-independent and where animal movements are being analyzed (Turchin, 1996). Geographic objects can be abstractly represented as lines, polygons, or points, and if the data are points, they may contain cartographic generalizations or location inaccuracy; therefore, scale is critical for analyzing point patterns (Wong and Lee, 2005). Boots and Getis (1988) discussed detecting spatial patterns in point distributions. The description of any spatial relationship between individual points requires the application of a special type of spatial statistics called centrographic measures (Kellerman, 1981). In geospatial analysis, if a location or the extent of a frame changes or if objects within a frame are repositioned, the results will also change (de Smith et al., 2007). Palancioğlu (2003) described these “movement signatures” resulting in movement patterns of objects using fuzzy and neuron-fuzzy approaches where uncertainties exist in the movement of the objects. Swarming is another approach that has been used in geospatial reasoning and is a method to explore possible paths through a landscape (Parunak et al., 2006). Central tendency measures in the spatial context will be the mean center used to define the centroid of a polygon assuming that all geographic features have the same weights or frequencies (Wise, 2002; Wong and Lee, 2005). The concept of average in classical statistics can be extended to the concept of geographic center, a measure of spatial central tendency, yet, there is no absolutely correct way to find the center of a spatial distribution (Wong and Lee, 2005). Standard distance is the spatial analogy of standard deviation in classical statistics and is expressed in distance units, which is a function of the coordinate system or projection adopted (Wong and Lee, 2005). In North America, the median center is more often defined as the center of minimum travel from all points (Wong and Lee, 2005). In its simplest form, systematic sampling selects regularly spaced locations to ensure complete coverage of the entire study and can be combined with random sampling such that the geographical space is divided systematically but sampling is random within each partitioned region (Wong and Lee, 2005). Chang et al. (2008) developed a spatio-temporal approach that identifies abnormal spatio-temporal clustering patterns. Geospatial or spatial analysis can be defined as the use of analytical methods that are robust and capable of operating over a range of spatial and temporal scales to describe geographic based datasets using spatial software (de Smith et al., 2007). Spatial analysis and spatial statistics are more than a half century old (Wong and Lee, 2005). Furthermore, spatial analysis is a rapidly changing field and increasingly, GIS packages are including analytical tools as standard built-in facilities or as optional toolsets, add-ins, or analysts (de Smith et al., 2007). Spatial data sets are becoming more prevalent and quantitative research by Senft et al. (1985) demonstrated yearling heifers used forage quality and quantity to select community preference using ad hoc regression models. However, regression models do not imply causal relationships between the variables being examined (Wong and Lee, 2005). Regression approaches have been used with points as well as site or grid square data; however, it may be difficult to reduce the effects of spatial correlation using this approach (Buckland and Elston, 1993). Krysl et al. (1989) noted that Latin square designs are infrequently used in range studies because of the potential row × column interactions that make treatment comparisons valid but tend to be conservative due to potential interactions. Partial least squares (Stone and Brooks, 1990) have been used and may be appropriate for behavior data because it allows modeling in situations where the number of explanatory variables is greater than the number of observations. Multivariate techniques often perform poorly or are not applicable to animal distribution data, especially categorical data such as that used to describe vegetation types (Stockwell and Noble, 1992). Covariance analysis can be used when land replication is not possible if the study has 3 or more stocking rates or grazing pressures (Riewe, 1961). Animal-to-animal variation is usually the greatest source of variation in grazing trials; hence, properly applied covariance analysis may be useful for identifying and controlling some sources of animal variation in grazing trials (Petersen and Lucas, 1960). When production per animal or production per unit area is the response variable of interest, the experimental unit is the paddock and the animals that forage on it (Bransby, 1989). Wang and Li (2005) have used stochastic process and differential equations to discuss the spatial and temporal patterns of foraging. Analysis of covariance allows you to readily compare if the relationship between grazing use and factors, such as distance to water, differ between treatments (see Bailey et al., 2006 and Bailey et al., 2008). Geospatial techniques (Renolen, 1999) are currently being used to track humans (Buliung and Remmel 2008), wildlife (Millspaugh and Marzluff, 2001), and domestic animals (Wade et al., 1998). Geospatial analyses are robust, and capable of operating over a range of spatial and temporal scales (de Smith et al., 2007). Spatial data are unique in that observations describing geographic phenomena or events have spatial dependency embedded in them as described in the First Law of Geography (Sui, 2004). At the center of all spatial analysis is the concept of place that requires a coordinate system to give a rigorous and precise definition of place or location (Guptill, 2001; Wong and Lee, 2005). Furthermore, geospatial data require visualization, so analyses must include the creation and manipulation of images on at least a 2D coordinate system, to produce maps, diagrams, charts, or more likely, surface analysis that involves 3D views of terrestrial activities and their associated tabular datasets (de Smith et al., 2007). Iliffe (1999) provided information on map projections. Because free-ranging livestock can move in 3D space, Renolen (1999) referred to them as “spatio-temporal” objects. Although cattle distribution has been treated as a transport theory or a first passage time problem (Cox and 65 external non-random environmental factors (Williams and Gillard, 1971). In rangeland conditions, patch or mosaic foraging may be even more pronounced than in mono- or di-cultures where dung fouling and location of water and minerals may inhibit uniform foraging and patch formation, and in turn, creates difficulty in assessing the forage availability (Coleman et al., 1989). Hill (2004) assembled a comprehensive assessment of remote sensing methods and applications for grasslands. methods to accurately understand the phenomena are under investigation (Unwin, 2009). The text books by Wong and Lee (2005) and de Smith et al. (2007) provide basic statistical concepts along with ready-to-use tools and programs in spatial analysis and statistics that are integrated with highly accessible GIS packages. Table 4 lists 46 different mathematical approaches that have been used to evaluate animal distribution and its various components. Austin (2007) suggested that the combination of ecological knowledge and statistical skill is more important than the precise statistical method used to evaluate species distribution, especially for conservation and climate change management. Most software packages have been developed for commercial use rather than research, making it difficult to use “canned” programs to manage data and develop new statistical methods for evaluating animal trajectories (Calenge, 2006; Calenge et al. 2009). Fotheringham et al. (2002) addressed the geographically weighted animal distribution results based on animal movement that consist of several parts, including animal trajectories that have recently been evaluated by Calenge et al. (2009). Tortuosity of an animal‟s trajectory may provide information on the animal‟s foraging strategies (Benhamou, 2004). The tortuosity of an animal‟s trajectory tends to reflect the intensity of the animal‟s search for food and for this reason has been studied in great detail (Calenge, 2009). An animal‟s use of its habitat is determined by its trajectory (Aebischer et al., 1993), and the number of animals used to obtain accurate data is critical. Aebischer et al. (1993) suggested 6 radio-tagged animals constitute an absolute minimum number when evaluating habitat use and this number should definitely be above 10, and preferably above 30, if attempting to describe a population since individual behavior varies. Thus, pooling data across animals is justifiable only if they do not differ. Wong and Lee (2005) indicated that the first effort to integrate spatial analysis and statistical functions with GIS began in 1999. Geospatial analysis is concerned with where something happens and makes use of geographic information that links features and phenomena on the Earth‟s surface to their locations (de Smith et al., 2007). Geospatial analysis reflects the melding of the broad field of spatial analysis (Cressie, 1991) with the latest generation of Geographic Information System (GIS; Longley et al., 2005) and related software (de Smith et al., 2007). An interactive text version on geospatial analysis is available on the web (www.spatialanalysisonline.com). When certain a phenomenon occurs, it may be due to a random process or a systematic process, and if part of a systematic process, the numeric or spatial patterns will be of interest (Wong and Lee, 2005). Procedures have been developed, based on random processes, to allow us to determine if a sample is a reliable representative of the population (Wong and Lee, 2005). Griffith and Amrhein (1991, pp 215) presented a sampling scheme that was designed to accommodate the sampling of observations in the geographic space referred to as spatial sampling (Wong and Lee, 2005). Pattern analysis can be complementary to statistical analysis of grazing experiments, especially if influenced by MODELS Models provide a logical framework in which to evaluate factors that impact a process. Modeling real world phenomena in space and time is a non-trivial task (Renolen, 1999). Basically all wildlife habitat evaluation models have attempted to evaluate the carrying capacity of the wildlife habitat for a particular wildlife species (Kushwaha and Roy, 2002). Van Dyne et al. (1980) listed 75 models that involve large herbivore foraging; only 1 was built to handle feeding behavior. There is no standard or current best practice when modeling a species environmental niche or geographical distribution, whether plant or animal Austin (2007). Certain models, such as the Hierarchical Object Orientated Foraging Simulator (HOOFS) have been used for experimentation with individual-oriented, spatially-explicit herbivore vegetation systems (Beecham and Farnsworth, 1998). Recent models have been developed that will detect areas of over- as well as underutilization using Landsat-TM imagery (Röder et al., 2007). Although Wiegand et al. (2008) provided a historical development in ecological theory to describe grazing models that stimulated and mirrored the conceptual changes and advances in the understanding of grazing systems, no model will ever allow one to take into account the entire complexity of animal behavior, and no universal recipe is available to analyze animals‟ trajectories (Calenge et al., 2009). Models are being developed using GIS data that integrate cattle nutritional requirements with satellite databased standing crop data to estimate foraging capacity on small (1 to 20 ha) paddocks (Phillips et al., 2009); however, animal behavior may not be simultaneously considered in these models. Furthermore, not all models appear to predict the distribution of foraging animals. Kennedy and Gray (1993) found the Ideal Free Distribution (IFD) model consistently under matched the distribution of resources with the distribution of organisms, suggesting that to accurately predict the foraging distribution of animals, much more information is required than just the relative distribution of resources. Marion et al. (2005) developed a model involving the interaction of the sense of sight and smell that demonstrated the importance of individual animal behaviors. In this model, sight attracted an animal to a particular location based on standing crop height and subsequently smell determined forage consumption based on the presence or absence of fecal contamination. Movement models have been constructed in an attempt to describe animal movement (Farnsworth and Beecham, 1999; Codling et al., 2008). In general, models that involve information garnered from field observations may provide a 66 grazing that uses a simple general formula for predicting sustainable levels of grazing if local site productivity is known; however, it lacked the ability to evaluate the effects of trampling and uneven grazing distribution. Regression modeling rarely examines correlation of variables in terms of a process (Austin, 2007). The Geographically Weighted Regression (GWR) model was developed to accommodate non-stationary parameters in a regression framework (Fotheringham et al., 2002). Techniques that address stationary parameters in spatial regressions have been addressed by Griffith (1988; 2003). Although random walk models suggest randomness, these models can be made very complex and nonrandom, being limited only by the capacity of computers to simulate and the capacity of the human brain to assimilate the results (Turchin, 1996). Using and combining more than a single method to describe foraging of free-ranging animals appears useful. Guo et al. (2009) developed a hierarchical modeling methodology that successfully combined Hidden Markov Models (HMM) and long-term prediction to describe the behavior among 6 cows foraging on 7 ha of tame pasture in Queensland, Australia. Doug Johnson at Oregon State University and a team have developed a computerized multifactor decision making tool that incorporates landscape parameters to determine the suitability of each site on the landscape in terms of animals usage (KRESS, 2006). More recently, Horne et al. (2008) developed a model that simultaneously couples home range behavior and resource selection analysis into a single model that links statistical and process-based movement approaches to home range modeling. Although patch utilization models appear to work with grazing herbivores, mounting evidence suggests applying patch optimization models to browsers may be incorrect (Searle and Shipley, 2008). richer set of biological detail than models that attempt to describe individual space-use patterns without this information (Don and Rennolls, 1983). For those that deal with livestock herd models, the question always arises concerning the degree of agreement between the model and the actual system being modeled (Sørensen, 1990). Most models implicitly assume homogenous or, at most, very simple heterogeneous environments in which no interaction between individuals is assumed (Codling et al., 2008). This is simply not the case for domestic livestock confronted with complicated environments that affect movement behavior (Vuilleumier and Metzger, 2006). Furthermore, with gregarious (i.e., herd or flocking) domestic animal species, interactions between and among individuals exist and these interactions have an important effect on overall animal dispersal (Couzin et al., 2005). Models have been developed to examine not only group behaviors (Stillman, 2008) but also behavior between animal pairs (Rands et al., 2008). Okubo (1980) modeled the spread of animals as a diffusion process using the suite of geostatistical methods known as kriging (Matheron, 1965). Kriging was the name given by Prof. Georges Matheron in honor of the South African mining engineer Danie Krig (Krige, 1951) who developed an optimal spatial regression geospatial interpolation technique for making spatial predictions at unobserved locations using point data from observed locations (Bailey and Gatrell, 1995; Stein, 1999). Cressie and Kornak (2002) derived a trend-estimation and kriging methodology for geostatistics in the presence of location error. Kriging has been used to describe animal distributions because of its ability to make predictions of processes in space while taking into account spatially correlated observations. However, kriging has the disadvantage that it is designed for point data rather than site or grid data. Cokriging refers to a technique used with spatially sampled curves, such as the dives of elephant seals (Nerini et al., 2010). Because many factors affect distribution, it should be recognized that as factors included in a model increase the complexity of the model, a concomitant reduction in model reliability occurs (Wang and Li, 2005). Horne and Garton (2006) suggested that using multiple models may be more appropriate than using a single model. Horne et al. (2008) suggested that fitting more than a single model to data is appropriate in order to reflect all the different assumptions and (or) hypotheses concerning the individual‟s use of space, and models should vary in complexity to guard against over- or under-fitting models. Hybrid models exist that utilize not only a motion function but also movement patterns useful in forecasting an objects future location (Jeung et al., 2008). Farnsworth and Beecham (1999) proposed a mathematical model that combines features of cognitive and non-cognitive navigation and unites foraging behaviors at different scales in order to model a wide range of behaviors. Regarding use of space (home-range), Trail and Bigalke (2006) used a GIS-based model, using presence–only data collected with a GPS receiver, to predict habitat suitability for large grazing ungulates. They found that greenness of grass was the main determinant of habitat utilization. Read et al. (2002) described a simple process-based model for HOME RANGE The term home range was coined by Burt (1943), and although it is usually associated with wildlife, the term can also be used with livestock. Furthermore, home range is defined without reference to the presence or absence of particular types of behavior (Brown and Orians, 1970). It refers to the area in which an individual conducts its normal activities of food gathering, mating, and caring for young. Burt (1943) makes a distinction between home range and territory in that territory is an area an animal will protect by fighting or display of aggressive gestures towards others of its own kind. In 1988, Reiss wrote “we still do not know the precise reasons why animals have the home range sizes they do.” Roath and Krueger (1982) suggested cattle home ranges persist from year to year. Using both GPS and GIS data, the “kernel home-range” method has been used to evaluate cattle grazing (Barbari et al., 2006). Home ranges (areas, rarely found in convenient geometric designs) are not static over time and the home ranges of different individuals may and do overlap and these areas tend to vary with sex, age, animal density, and season (Burt, 1943). Loehle (1990) noted that fractals may be an appropriate way to investigate home range since within a home range an animal will use some areas repeatedly while other areas 67 involves gathering data focused on animal distribution and the second involves appropriately analyzing it. Unfortunately, the lack of theory and the lack of data appear to be mutually reinforcing (Grünbaum, 1998). Global positioning system technology currently seems to be the preferred tool of choice to study the spatial and temporal patterns of landscape use by free-ranging animals. Fully understanding the geographical linkage of GPS animal location data within rangeland landscapes will require range animal ecologists to work as members of research teams composed of a number of different disciplines, including statisticians and modelers trained in geospatial mathematics. Based on the literature perused for this paper, there does not seem to be a single protocol that combines GPS and GIS into a “single best” package for evaluating all situations involving animal distribution. The wildlife profession has led in much of the foundational research on animal distribution. The range-animal livestock ecologist must adopt or modify existing approaches using discernment while developing new tools, techniques, and protocols where those in wildlife either do not exist or are inappropriate. Banhazi and Black (2009) suggested that the greatest technological progress in precision livestock farming will come from developing new technological tools rather than fine-tuning existing methods. Furthermore, the most exciting application of a new technology may not come from its intended application but when it is applied outside its original field (Rogers, 2001), and here is where assembling teams with the required experience and expertise can provide their greatest benefit. Current precision agriculture methodologies are focused on arable land rather than grassland (Schellberg et al., 2008). This will change as sensor technology advances concurrently with our understanding of how free-ranging animals use space and how best to manage the plant-animal interface. One broad area of research that is needed immediately involves determining which animals and how many should be assembled to form stable spatial and temporal freeranging groups designed to accomplish a specific goal for a particular landscape and configuration of paddock(s) and associated working facilities. The future will encounter larger and larger volumes of data that must be processed, summarized, and analyzed thus forming ever growing linkages of animal data with GIS data from landscapes, producing increasingly complex geospatial models (Kenward, 2001b). As this occurs, research specifically addressing optimum geospatial protocols for applying statistical techniques to evaluate free-ranging livestock distribution seems to be a fruitful area to parallel biological investigations into free-ranging animal behavior. are never used and all this kind of detail is lost if a line is simply drawn around the points where the animal has been seen. Might fractals be used to calculate more accurate stocking rates? Both parametric and nonparametric methods can be used to investigate home-range data, but nonparametric approaches have the great advantage of flexibility (Worton, 1989). Worton (1987) has described and compared various home range models and considers the kernel method and Fourier Transform method to provide good flexibility. Autocorrelation should also be considered in radio-tracking data especially those used for home range estimation (Otis and White, 1999). Kernel-density estimation (KDE) is one of the most widely used home range estimators in ecology (Hemson et al., 2005); however, results of Hemson et al. (2005) indicated that it may not be the most reliable method for estimating home range. Kernel-based estimates from small samples are poor for identifying fine structure and will over-estimate the home range size (Seaman and Powell, 1996). The disadvantage of kernel analysis lies in its failure to recognize the asymmetry of animal movement patterns, and “canned” programs for kernel utilization distribution (UD) estimation do not account for this asymmetry (Amstrup et al., 2004). Kernel methods consist of placing a kernel (a probability density) over each observation point in the sample and if kernel-based estimates are derived from small samples, the results will overestimate home range size (Seaman et al., 1999). Therefore, Seaman et al. (1999) recommended that home range studies using kernel estimators should have representative samples from over 30 locations and preferably more than 50. Because of the importance of sample size, it was strongly recommended that authors report sample sizes in addition to the exact home range estimation method used. Comparing home range size estimates among researchers can lead to inaccurate conclusions (GalleraniLawson and Rodgers, 1997). Those authors found that using the same data to calculate home-range sizes with 5 commonly used home-range programs to determine minimum convex polygon, harmonic mean, and kernel estimators at 3 levels of resolution (95%, 75%, and 50% of locations) produced different results. They further suggested research manuscripts include in their materials and methods section the home range estimator software used, the reason it was selected, the user-selected options, and input values of required parameters. CONCLUSIONS Monitoring free-ranging animals to accurately quantify and subsequently understand distribution is not a trivial task. Asking the proper researchable questions in the proper sequence, using the correct equipment to collect data, and finally analyzing data using the most appropriate geospatial statistical tools are the essential steps required to reach accurate conclusions. Although the list of factors currently known to affect animal distribution is lengthy, most scientists would agree the list is incomplete and more research is warranted (Bailey, 2005). The type of research needed can be divided into two broad areas. The first SPECIAL CONSIDERATIONS The literature reviewed in this paper has revealed several points that bear consideration when conducting research into animal distribution. First, there remains a need to focus on understanding how abiotic as well as internal and external biotic factors drive group fission and fusion and the subsequent leadership within groups. Second, experimental designs should attempt to embrace how the animal perceives its world rather than how the researcher 68 American Society of Agronomy, American Dairy Science Association, American Society of Animal Production, and American Society of Range Management. Joint Committee. 1952. Pasture and range research techniques. Agron. J. 44:39-50. Amlaner, C. J., Jr. and D. W. Macdonald (eds). 1980. A Handbook on Biotelemetry and Radio Tracking. Pergamon Press. Oxford, UK. Amstrup, S. C., T. L. McDonald, and G. M. Durner. 2004. Using satellite radiotelemetry data to delineate and manage wildlife populations. Wildl. Soc. Bul. 32:661679. Anderson, D. M. 1988. Seasonal stocking of tobosa managed under continuous and rotational grazing. J. Range Manage. 41:78-83. Anderson, D. M. 1998. Pro-active livestock management – capitalizing on animal behavior. J. Arid Land Stud. 7S:113-116. Anderson, D. M. 2007. Virtual fencing – past, present and future. Rangeland J. 29:65-78. Anderson, D. M. and J. L. Holechek. 1983. Diets obtained from esophageally fistulated heifers and steers simultaneously grazing semidesert tobosa rangeland. Proc. West Sec. Soc. Anim. Sci. 34:161-164. Anderson, D. M. and N. S. Urquhart. 1986. Using digital pedometers to monitor travel of cows grazing arid rangeland. Appl. Anim. Behav. Sci. 16:11-23. Anderson, D. M., C. V. Hulet, J. N. Smith, W. L. Shupe, and L. W. Murray. 1987. Heifer disposition and bonding of lambs to heifers. Appl. Anim. Behav. Sci. 19:27-30. Anderson, D.M., C.V. Hulet, W.L. Shupe, J.N. Smith, and L.W. Murray. 1988. Response of bonded and nonbonded sheep to the approach of a trained border collie. Appl. Anim. Behav. Sci. 21:251-257. Arabie, P., L. J. Hubert, and G. D. Soete (eds). 1996. Clustering and Classification. World Scientific Publ. River Edge, NJ. Archibald, S. and W. J. Bond. 2004. Grazer movements: spatial and temporal responses to burning in a tall-grass African savanna. Int. J. Wildland Fire. 13:377-385. Arnold, G. W., and M. L. Dudzinski. 1978. Ethology of Free-Ranging Domestic Animals. Elsevier Scientific Publishing, New York, NY. Austin, M. 2007. Species distribution models and ecological theory: A critical assessment and some possible new approaches. Ecol. Model. 200:1-19. Barbari, M., L. Conti, B. K. Koostra, G. Masi, F. S. Guerri, and S. R. Workman. 2006. The use of global positioning and geographical information systems in the management of extensive cattle grazing. Biosyst. Eng. 95:271-280. Bailey, D. W. 1988. Characteristics of spatial memory and foraging behavior in cattle. Ph.D. Diss., Colorado State Univ., Fort Collins. Bailey, D.W. 1995. Daily selection of feeding areas by cattle in homogeneous and heterogeneous environments. Appl. Anim. Behav. Sci. 45:183-200. Bailey, D. W. 2004. Management strategies for optimal grazing distribution and use of arid rangelands. J. Anim. Sci. 82:E147-153. perceives the world. Third, outliers normally considered a nemesis to most scientists attempting to compare treatment means may provide more useful information to understanding free-ranging livestock behavior than does the mean. If this is true, new and innovative statistical tools will need to be developed to analyze behavioral data. Forth, because behavior among free-ranging individuals can be quite variable, behavioral studies should be conducted using larger animal numbers than considered minimal for classical animal science nutrition research. Fifth, because size and composition of animal groupings affect behaviors, applying research results should be done conservatively unless conditions are “identical”. Sixth, statistics should only serve as a suggestion for the presence or absence of behaviors that may have cause and effect relationships. Seventh, livestock behavior has been influenced by man‟s selection pressures; therefore, extrapolating wildlife behaviors to domestic animals should only be done with discernment. Eight, how best to collect and analyze freeranging animal behavioral data aimed at understanding animal distribution is not a trivial undertaking; therefore, developing workshops focused on bringing together the various disciplines required to address the most pressing questions concerning free-ranging animal behavior seems appropriate. Workshop topics should address the following questions. What part of animal behavior studies can and should be replicated? What experimental designs should be used to examine the spatial and temporal formation of groups? What geospatial tools provide the most accurate results for predicting distribution behaviors? What part of free-ranging animal behavior studies should be standardized to make results more universally applicable? LITERATURE CITED Adams, D. C. 1985. Effect of time of supplementation on performance, forage intake and grazing behavior of yearling beef steers grazing Russian wild ryegrass in the fall. J. Anim. Sci. 61:1037-1042. Aebischer, N. J., P. A. Robertson, and R. E. Kenward. 1993. Compositional analysis of habitat use from animal radio-tracking data. Ecol. 74:1313-1325. Agreil, C. and M. Meuret. 2004. An improved method for quantifying intake rate and ingestive behaviour of ruminants in diverse and variable habitats using direct observation. Small Rumin. Res. 54:99-113. Aharoni, Y., Z. Henkin, A. Ezra, A. Dolev, A. Shabtay, A. Orlov, Y. Yehuda, and A. Brosh. 2009. Grazing behavior and energy costs of activity: A comparison between two types of cattle. J. Anim. Sci. 87:2719-2731. Aitchison, J. 1986. The Statistical Analysis of Compositional Data. Chapman and Hall. London, England. Allan, A. L. 2004. Maths for Map Makers. 2nd ed. Whittles Publishing, Dubeath, Scotland. Allden, W. G. 1962. Rate of herbage intake and grazing time in relation to herbage availability. Proc. Aust. Soc. Anim. Prod. 4:1163-1171. 69 Bluman, A. G. 2003. Elementary Statistics: A Step By Step Approach. 5th ed. McGraw Hill. New York, NY. Boehlert, G. W., D. P. Costa, D. E. Crocker, P. Green, T. O'Brien, S. Levitus, and B. J. Le Bouef. 2001. Autonomous pinniped environmental samplers: using instrumented animals as oceanographic data collectors. J. Atmos. Ocean Tech. 18:1882-1893. Bolstad, P., A. Jenks, J. Berkin, K. Horne, and W.H. Reading. 2005. A comparison of autonomous, WAAS, real-time, and post-processed global positioning systems (GPS) accuracies in northern forests. Northern J. Appl. Forestry. 22:5-11. Bond, J., G.E. Carlson, C. Jackson Jr., and W.A. Curry. 1967. Social cohesion of steers and sheep as a possible variable in grazing studies. Agron. J. 59:481-482. Bonham, C. D. (convenor). 1986. Statistical analyses and modeling of grazing systems. Symposium Proceedings. February 11, 1986, Kissimmee, FL. Soc. Range Manage. Denver, CO. Boots, B. N. 1980. Weighting Thiessen polygons. Econ. Geog. 56:248-259. Boots, B. N. and A. Getis. 1988. Point Pattern Analysis. Sage Publications. Newberry Park, CA. Börger, L., B. D. Dalziel, and J. M. Fryxell. 2008. Are there general mechanisms of animal home range behaviour? A review and prospects for future research. Ecol. Letters. 11:637-650. Box, T. W., G. W. Brown, and J. Liles. 1965. Influence of winter supplemental feeding of cottonseed cake on activities of beef cows. J. Range Manage. 18:124-126. Bransby D. I. 1989. Compromise in the design and conduct of grazing experiments. Pages 53-67 In Grazing Research: Design, Methodology, and Analysis. G. C. Marten (ed) CSSA Spec. Pub. No. 16. Crop Sci. Soc. Amer., Madison, WI. Breiman, L., J. H. Friedman, R. A. Olshen, and C. J. Stone. 1984. Classification And Regression Trees. Wadsworth. Belmont, CA. Brock, B. L. and C. E. Owensby. 2000. Predictive models for grazing distribution: A GIS approach. J. Range Manage. 53:39-46. Brosh, A., Z. Henkin, E. D. Ungar, A. Dolev, A. Shabtay, A. Orlov, Y. Yehuda, and Y. Ahroni. 2010. Energy cost of activities and locomotion of grazing cows: A repeated study in larger plots. J. Anim. Sci. 88:315-323. Brown, J. L. and G. H. Orians. 1970. Spacing patterns in mobile animals. Ann. Rev. Ecol. Sust. 1:239-262. Brunson, M. W. and E. A. Burritt. 2009. Behavioral factors in rotational grazing systems. Rangelands 31:20-25. Buckland, S. T. and D. A. Elston. 1993. Empirical models for the spatial distribution of wildlife. J. Appl. Ecol. 30:478-495. Buerkert, A. and E. Schlecht. 2009. Performance of three GPS collars to monitor goats‟ grazing itineraries on mountain pastures. Comp. Elect. Agric. 65:85-92. Buliung, R. N. and T. K. Remmel. 2008. Open source, spatial analysis, and activity-travel behaviour research: Capabilities of the aspace package. J. Geog. Syst. 10:191-216. Bailey, D. W. 2005. Identification and creation of optimum habitat conditions for livestock. Rangeland Ecol. Manage. 58:109-118. Bailey, D. W. and L. R. Rittenhouse. 1989. Management of cattle distribution. Rangelands. 11:159-161. Bailey, D. W., J. E. Gross, E. A. Laca, L. R. Rittenhouse, M. B. Coughenour, D. W. Swift, and P. L. Sims. 1996. Mechanisms that result in large herbivore distribution patterns. J. Range Manage. 49:386-401. Bailey, D. W., J. W. Walker, and L. R. Rittenhouse. 1990. Sequential analysis of cattle location: Day-to-day movement patterns. Appl. Anim. Behav. Sci. 25:137148. Bailey, D. W., G. R. Welling, and E. T. Miller. 2001. Cattle use of foothills rangeland near dehydrated molasses supplement. J. Range Manage. 54:338-347. Bailey, D. W., H. C. VanWagoner, and R. Weinmeister. 2006. Individual animal selection has the potential to improve uniformity of grazing on foothill rangeland. Rangeland Ecol. Manage. 59:351-358. Bailey, D. W., H. C. VanWagoner, R. Weinmeister, and D. Jensen. 2008. Comparison of low-moisture blocks and salt for manipulating grazing patterns of beef cows. J. Anim. Sci. 86:1271-1277. Bailey, T. C. and A. C. Gatrell. 1995. Interactive Spatial Data Analysis. Longman. Essex, UK. Bajaj, R., S. L. Ranaweera, and D. P. Agrawal. 2002. GPS: location-tracking technology. Computer 35:92-94. Banhazi, T. M. and J. L. Black. 2009. Precision livestock farming: A suite of electronic systems to ensure the application of best practice management on livestock farms. Aust. J. Multi-discipl. Eng. 7:1-14. Barnes, M. K., B. E. Norton, M. Maeno, and J. C. Malechek. 2008. Paddock size and stocking density affect spatial heterogeneity of grazing. Rangeland Ecol. Manage. 61:380-388. Barton, R. K., L. J. Krysl, M. B. Judkins, D. W. Holocombe, J. T. Broesder, S. A. Gunter, and S. W. Beam. 1992. Time of daily supplementation for steers grazing dormant intermediate wheatgrass pasture. J. Anim. Sci. 70:547-558. Baumont, R., S. Prache, M. Meuret, and P. Morand-Fehr. 2000. How forage characteristics influence behaviour and intake in small ruminants: a review. Livest. Prod. Sci. 64:15-28. Beecham, J. A. and K. D. Farnsworth. 1998. Animal foraging from an individual perspective: An object orientated model. Ecol. Model. 113:141-156. Benhamou, S. 2004. How to reliably estimate the tortuosity of an animal‟s path: straightness, sinuosity, or fractal dimension? J. Theor. Biol. 229:209-220. Beringer, J., J. J. Millspaugh, J. Sartwell, and R. Woeck. 2004. Real-time video recording of food selection by captive white-tailed deer. Wildl. Soc. Bull. 32:648-654. Bharucha-Reid, A. T. 1960. Elements of The Theory of Markov Processes and Their Applications. McGrawHill. New York, NY. Blake, E. H., L. Steventon, J. Edge, and A Foster. 2002. A field computer for animal trackers. Pages 532-533 In Association for Computing Machinery Inc. CHI Conf. Pretoria, S. Afr. 70 Burghardt, T., and J. Ćalić. 2006. Analysing animal behaviour in wildlife videos using face detection and tracking. IEE Proc. Vis. Image Signal Process. 153:305312. Burns, J. C., H. Lippke, and D. S. Fisher. 1989. The Relationship of herbage mass and characteristics to animal responses in grazing experiments. Pages 7-20 In Grazing Research: Design, Methodology, and Analysis. G.C. Marten (ed) CSSA Spec. Pub. No. 16. Crop Sci. Soc. Amer., Madison, WI. Burns, J. C. and J. E. Standaert. 1985. Productivity and economics of legume-based vs. nitrogen-fertilized grassbased pastures in the United States. Pages 56-71 In Forage Legumes For Energy-Efficient Animal Production. R. F. Barnes, P. R. Ball, R. W. Brougham, G. C. Marten, and D. J. Minson (eds) Proc. Trilateral Workshop, Palmerston North, NZ April 30 - May 4, 1984. USDA-ARS, Washington DC. Butler, D. R. 2002. Visualizing animal impacts on the landscape: Remote sensing in the Geography classroom. Geocarto Int. 17:69-76. Burt, W. H. 1943. Territoriality and home range concepts as applied to mammals. J. Mamm. 24:346–352. Calenge, C. 2006. The package adehabitat for the R software: A tool for the analysis of space and habitat use by animals. Ecol. Model. 197:516-519. Calenge, C., S. Dray, and M. Royer-Carenzi. 2009. The concept of animals‟ trajectories from a data analysis perspective. Ecol. Informatics. 4:34-41. Canaway, R. J., W. F. Raymond, and J. C. Taylor. 1955. The automatic recording of animal behaviour in the field. Electr. Eng. 27:102-105. Casaer, J., M. Hermy, P. Coppin, and R. Verhagen. 1999. Analysing space use patterns by Thiessen polygon and triangulated irregular network interpolation: A nonparametric method for processing telemetric animal fixes. Int. J. Geog. Info. 13:499-511. Castle, M. E., A. S. Foot, and R. J. Halley. 1950. Some observations on the behaviour of dairy cattle with particular reference to grazing. J. Dairy Res. 17:215232. Chang, W., D. Zeng, and H. Chen. 2008. A stack-based prospective spatio-temporal data analysis approach. 2008. Dec. Support Syst. 45:697-713. Chapline, W. R. and M. W. Talbot. 1926. The use of salt in range management. U.S. Dept. Agric. Circ. 379. U.S. Govt. Printing Off. Washington DC. Chase, I. D. 1980. Cooperative and noncooperative behavior in animals. Amer. Nat. 115:827-857. Christman, M. C. and E. H. Leone. 2007. Statistical aspects of the analysis of group size effects in confined animals. Appl Anim. Behav. Sci. 103:265-283. Christman, M. C. and D. Lewis. 2005. Spatial distribution of dominant animals with a group: comparison of four statistical test of location. Anim. Behav. 70:73-82. Clark, P. E., D. E. Johnson, M. A. Kniep, P. Jermann, B. Huttash, A. Wood, M. Johnson, C. McGillivan, and K. Titus. 2006. An advanced, low cost, GPS-based animal tracking system. Rangeland Ecol. Manage. 59:334-340. Clary, W. P., P. F. Ffolliott, and F. R. Larson. 1978. Factors affecting foraging consumption by cattle in Arizona ponderosa pine forests. J. Range Manage. 31:9-13. Cliff, A. and J. K. Ord. 1981. Spatial Processes: Models and Applications. Pion. London, UK. Cochran, W. W., D. W. Warner, J. R. Tester, and V. B. Kuechle. 1965. Automatic radio-tracking system for monitoring animal movements. BioSci. 15:98-100. Coffey, C. 2009. Livestock distribution. http://www.noble.org/Ag/Forage/LivestockDistribution/ index.html. Accessed Oct. 17 2009. Coleman, S. W., T. D. A. Forbes, and J. W. Stuth. 1989. Measurements of the plant-animal interface in grazing research. Pages 37-52 In Grazing Research: Design, Methodology, and Analysis. G.C. Marten (ed) CSSA Spec. Pub. No. 16. Crop Sci. Soc. Amer. Madison, WI. Codling, E. A., M. J. Plank, and S. Benhamou. 2008. Random walk models in biology. J. R. Soc. Interface 5:813-834. Cook, C. W. and J. Stubbendieck. 1986. Range research: Basic problems and techniques. Soc. Range Manage. Denver, CO. Cooper, S. M., H. L. Perotto-Baldivieso, M. K. Owens, M. G. Meek, and M. Figueroa-Pagán. 2008. Agric. Ecosyst. Environ. 127:85-92. Cory, V. L. 1927. Activities of livestock on the range. Texas Agric. Exp. Stn. Bull. 367. 47 pp. Coughenour, M. B. 1991. Spatial components of plantherbivore interactions in pastoral, ranching, and native ungulate ecosystems. J. Range Manage. 44:530-542. Couzin, I. D., J. Krause, N. R. Franks, and S. A. Levin. 2005. Effective leadership and decision-making in animal groups on the move. Nature 433:513-516. Cox, S. 2002. Information technology: the global key to precision agriculture and sustainability. Comp. Elect. Agric. 36:93-111. Cox, D. R. and H. D. Miller. 1965. The Theory of Stochastic Processes. Wiley, New York, NY. Curtis, S. E. and K. A. Houpt. 1983. Animal ethology: its emergence in animal science. J. Anim. Sci. 57(Suppl 2):234-247. Cutler, T. L. and D. E. Swann. 1999. Using remote photography in wildlife ecology: a review. Wildl. Soc. Bull. 27:571-581. Cracknell, A. P. 1998. Synergy in remote sensing – what‟s in a pixel. Int. J. Remote Sens. 19:2025-2047. Cressie, N. A. C. 1991. Statistics for Spatial Data. Wiley. New York, NY. Cressie, N. and J. Kornak. 2002. Spatial statistics in the presence of location error with an application to remote sensing of the environment. Tech. Rep. No. 71. http://www.amstatonline.org/sections/envr/envrann/StatGIS/Cressie.pdf Accessed Dec. 5, 2009. Croney, C. C. and R. C. Newberry. 2007. Group size and cognitive processes. Appl. Anim. Behav. Sci. 103:215228. Dale, P. 2005. Introduction to Mathematical Techniques Used in GIS. CRC Press. Boca Raton, FL. 71 East, B. 2003. The art of range riding: A coat of many colors. Rangelands 25:9-15. Ehrenreich, J. H. and A. J. Bjugstad. 1966. Cattle grazing time is related to temperature and humidity. J. Range Manage. 19:141-142. Estevez, I, I. L. Andersen, and E. Naevdal. 2007. Group size, density and social dynamics in farm animals. Appl. Anim. Behav. Sci. 103:185-204. Estevez, N. and D. Tatem. 2007. GIS tutorial: How to use ArcMap 9.1. http://www.trincoll.edu/depts/cc/documentation/GIS/Ho wToArcMap.pdf Accessed Dec. 04, 2009. Farnsworth, K. D. and J. A. Beecham. 1999. How do grazers achieve their distribution? A continuum of models from random diffusion to the ideal free distribution using biased random walks. The Amer. Nat. 153:509-526. Ferreira, J. C. B. C. and J. Waldmann. 2007. Covariance intersection-based sensor fusion for sounding rocket tracking and impact area prediction. Control Eng. Pract. 15:389-409. Findholt, S. L., B. K. Johnson, L. L. McDonald, J. W. Kern, A. Ager, R. J. Stussy, and L. D. Bryant. 2002. Adjusting for radiotelemetry error to improve estimates of habitat use. U.S. Dept. Agric., For. Serv. Pac. Northwest Res. Stn. Gen. Tech. Rep. PNW-GTR-555. Portland, OR. Fleming, J. S. and J. P. Tracey. 2008. Preface to „Aerial surveys of wildlife: Theory and applications.‟ Wildl. Res. 35:iii-iv. Flynn, E. S., C. T. Dougherty, and O. Wendroth. 2008. Assessment of pasture biomass with the normalized difference vegetation index from active ground-based sensors. Agron. J. 100:114-121. Folks, J. L. and R. S. Chhikara. 1978. The inverse Gaussian distribution and its statistical application – a review. J. Royal Statist. Soc. Series B. 40:263-289. Forester, J. D., A. R. Ives, M. G. Turner, D. P. Anderson, D. Fortin, H. L. Beyer, D. W. Smith, and M. S. Boyce. 2007. State-space models link elk movement patterns to landscape characteristics in Yellowstone National Park. Ecol. Monogr. 77:285-299. Fotheringham, A. S., C. Brunsdon, and M. E. Charlton. 2002. Geographically weighted regression: The analysis of spatially varying relationships. John Wiley & Sons. Chichester, UK. Fortin, D., M-E. Fortin, H.L. Beyer, T. Duchesne, S. Courant, and K. Dancose. 2009. Group-size-mediated habitat selection and group fusion-fission dynamics of bison under predation risk. Ecol. 90:2480-2490. Franklin, D. H., M. L. Cabrera, H. L. Byers, M. K. Matthews, J.G. Andrae, D. E. Radcliffe, M. A. McCann, H. A. Kuykendall, C. S. Hoveland, and V. H. Calvert, II. 2009. Impact of water troughs on cattle use of riparian zones in the Georgia Piedmont in the United States. J. Anim. Sci. 87:2151-2159. Frost, A. R., C. P. Schofield, S. A. Beaulah, T. T. Mottram, J. A. Lines, and C. M. Wathes. 1997. A review of livestock monitoring and the need for integrated systems. Comp. Elect. Agric. 17:139-159. Dark, S. J., and D. Bram. 2007. The modifiable areal unit problem (MAUP) in physical geography. Prog. Phy. Geo. 31:471-479. Darr, M. and W. Epperson. 2009. Embedded sensor technology for real time determination of animal lying time. Comp. Elect. Agric. 66:106-111. Davies, W. 1961. Research Techniques in Use at the Grassland Research Institute Hurley. Lamport Gilbert and Co., Ltd. Reading, UK. DeCesare, N. J., J. R. Squires, and J. A. Kolbe. 2005. Effect of forest canopy on GPS-based movement data. Wildl. Soc. Bull. 333:935-941. D‟Eon, R. and D. Delparte. 2005. Effects of radio-collar position and orientation on GPS radio-collar performance, and the implication of DPDOP in data screening. J. Appl. Ecol. 42:383-388. Dean, R. E. and R. W. Rice. 1974. Effects of fences and corrals on grazing behavior. Proc. West. Sec. Amer. Soc. Anim. Sci. 25:56-57. Deering, D. W., J. W. Rouse Jr., R. H. Haas, and J. A. Schell. 1975. Measuring forage production of grazing units from Landsat MSS data. Pages 1169-1178 In 10th Int. Symp. Remote Sens. Environ. Ann Arbor, MI. de Smith, M. J. 2006. Maths for the Mystified. Troubador Publishing Ltd. Leicester, UK. de Smith, M. J., M. F. Goodchild, and P. A. Longley. 2007. Geospatial Analysis A Comprehensive Guide to Principles, Techniques and Software Tools. The Winchelsea Press. Leicester, UK. DelCurto, T., M. Porath, C. T. Parsons, and J. A. Morrison. 2005. Management strategies for sustainable beef cattle grazing on forested rangelands in the Pacific Northwest. Rangeland Ecol. Manage. 58:119-127. Dixon, K. R. and J. A. Chapman. 1980. Harmonic mean measure of animal activity areas. Ecol. 61:1040-1044. Dolev, A., E. D. Ungar, and Z. Henkin. 2008. Factors influencing the spatial distribution of beef cattle on Mediterranean rangeland. Page 661 In Multifunctional Grasslands in a Changing World Vol. 1. Organizing committee (eds) IGC/IRC., Hohhot, China June 29-July 5. Guangdong People‟s Pub. House, Beijing, China. Don, B. A. C. and K. Rennolls. 1983. A home range model incorporating biological attraction points. J. Anim. Ecol. 52:69-81. Dudzinski, M. L., W. J. Müller, W. A. Low, and H. J. Schuh. 1982. Relationship between dispersion behaviour of cattle and forage conditions. Appl. Anim. Ethol. 8:225-241. Dumont. B., P. Carrère, and P. D‟Hour. 2002. Foraging in patchy grasslands: diet selection by sheep and cattle is affected by the abundance and spatial distribution of preferred species. Anim. Res. 51:367-381. Dunbar, R. I. M. 1992. Neocortex size as a constraint on group size in primates. J. Human Evol. 22:469-493. Dussault, C., R. Courtois, J. P. Ouellet, and J. Huot. 2000. Evaluation of GPS telemetry collar performance for habitat studies. Wildl. Soc. Bull. 27:965-972. Dwyer, D. L. 1961. Activities and grazing preferences of cows with calves in Northern Osage County, Oklahoma. Okla. Agric. Exp. Stn. Bull. 588. 61pp. 72 Gregorini, P., S. A. Gunter, P. A. Beck, J. Caldwell, M. T. Bowman, and W. K. Coblentz. 2009. Short-term foraging dynamics of cattle grazing swards with different canopy structures. J. Anim.Sci. 87:3817-3824. Griffith, D. A. 1987. Toward a theory of spatial statistics: Another step forward. Geograph. Anal. 19:69-82. Griffith, D. A. 1988. Advanced Spatial Statistics. Kluwer Academic. Boston, MA. Griffith, D. A. 2003. Spatial Autocorrelation and Spatial Filtering. Springer-Verlag. Germany, Berlin. Griffith, D. A. 2009. Spatial autocorrelation. Pages 1-10 in The International Encyclopedia of Human Geography. R. Kitchin and N. Thrift (eds) Elsevier Inc., Amsterdam, The Netherlands. Griffith, D.A. and C. G. Amrhein. 1991. Statistical analysis for geographers. Prentice Hall. Englewood Cliff, N.J. Griffiths, W. M., V. Alchanatis, R. Nitzan, V. Ostrovsky, E. Ben-Moshe, R. Yonatan, S. Brener, H. Baram, A. Genizi, and E. D. Ungar. 2006. A video and acoustic methodology to map bite placement at the patch scale. Appl. Anim. Behav. Sci. 98:196-215. Grünbaum, D. 1998. Using spatially explicit models to characterize foraging performance in heterogeneous landscapes. Amer. Nat. 151:97-115. Gueron, S., S. A. Levin, and D. I. Rubenstein. 1996. The dynamics of herds: From individuals to aggregations. J. Theor. Biol. 182:85-98. Guo, Y., G. Poulton, P. Corke, G. J. Bishop-Hurley, T. Wark, and D. L. Swain. 2009. Using accelerometer, high sample rate GPS and magnetometer data to develop a cattle movement and behaviour model. Ecol. Model. 220:2068-2075. Guptil, S. C. 2001. Spatial data. Pages 14775-14778 In International Encyclopedia of the Social and Behavioral Sciences. N. J. Smelser and P.B. Baltes (eds) Elsevier, Maryland Heights, MO. Hacker, R. B., B. E. Norton, M. K. Owens, and D. O. Frye. 1988. Grazing of crested wheatgrass, with particular reference to effects of pasture size. J. Range Manage. 41:73-78. Hafez, E. S. E. 1968. Behavioral adaptations. Pages 202214 In Adaptation of Domestic Animals. E.S.E. Hafez (ed) Lea and Febiger. Baltimore, MD. Hagen, C. A., N. C. Kenkel, D. J. Walker, R. K. Baydack, and C. E. Braun. 2001. Fractal-based spatial analysis of radiotelemetry data. Page 167-187 In Radio Tracking And Animal Populations. J. J. Millspaugh and J. M. Marzluff (eds) Academic Press, San Diego, CA. Haining, R. 2003. Spatial Data Analysis – Theory And Practice. Cambridge Univ. Press. Cambridge, UK. Hancock, J. 1953. Grazing behavior of cattle. Anim. Bred. Abstr. 21:1-13. Hargrove, W. W. and J. Pickering. 1992. Pseudoreplication: a sine qua non for regional ecology. Landscape Ecol. 6:251-258. Harris, N. R., D. E. Johnson, M. R. George, and N. K. McDougald. 2002. The effect of topography, vegetation, and weather on cattle distribution at the SanJoaquin Experimental Range, California. Pages 53-56 In USDA For. Serv. Gen. Tech. Rep. PSW-GTR-184. USDA Forest Service, Albany, CA. Gallerani-Lawson, E. J. and A. R. Rodgers. 1997. Differences in home-range size computed in commonly used software programs. Wildl. Soc. Bull. 25:721-729. Ganskopp, D. 2001. Manipulating cattle distribution with salt and water in large arid-land pastures: A GPS/GIS assessment. Appl. Anim. Behav. Sci. 73:251-262. Ganskopp, D., R. Cruz, and D. E. Johnson. 2000. Leasteffort pathways? A GIS analysis of livestock trails in rugged terrain. Appl. Anim. Behav. Sci. 68:179-190. Ganskopp, D. and D. Bohnert. 2006. Do pasture-scale nutritional patterns affect cattle distribution on rangelands? Rangeland Ecol. Manage. 59:189-196. Ganskopp, D. C. and D. W. Bohnert. 2009. Landscape nutritional patterns and cattle distribution in rangeland pastures. Appl. Anim. Behav. Sci. 116:110-119. Garcia, F., P. Carrère, J. F. Soussana, and R. Baumont. 2005. Characterisation by fractal analysis of foraging paths of ewes grazing heterogeneous swards. Appl. Anim. Behav. Sci. 93:19-37. Gary, L. A., C. W. Sherritt, and E. B. Hale. 1970. Behavior of Charolais cattle on pasture. J Anim. Sci. 30:203-206. Geographic Resource Analysis Support System (GRASS). 2000. A raster, vector Geographical Information System (GIS) software program Ver. 5.0beta6. http://www.boutell.com/lsm/lsmbyid.cgi/000876 Accessed Dec. 01, 2009. Gerard, J-F., E. Bideau, M-L. Maublanc, P. Loisel, and C. Marchal. 2002. Herd size in large herbivores: Encoded in the individual or emergent? Biol. Bull. 202:275-282. Gibb, M. J., C. A. Huckle, R. Nuthall, and A. J. Rook. 1999. The effect of physiological state (lactating or dry) and sward surface height on grazing behaviour and intake by dairy cows. Appl. Anim. Behav. Sci. 63:269287. Gillespie, T. W. 2001. Remote sensing of animals. Prog. Phys. Geog. 25:355-362. Gilsdorf, J. M., K. C. Vercauteren, S. E. Hygnstrom, W. D. Walter, J. R. Boner, and G. M. Clements. 2008. An integrated vehicle-mounted telemetry system for VHF telemetry applications. J. Wildl. Manage. 72:1241-1246. Girard, I., J. P. Ouellet, R. Courtois, C. Dussault, and L. Breton. 2002. Sampling effort based on GPS telemetry on home-range size estimations. J. Wildl. Manage. 66:1290-1300. Girard, I., C. Dussault, J. P. Ouellet, and R. Courtois. 2006. Balancing number of locations with number of individuals in telemetry studies. J. Wildl. Manage. 70:1249-1256. Gordon, I. J. 1995. Animal-based techniques for grazing ecology research. Small Rumin. Res. 16:203-214. Graetz, R. D. 1987. Satellite remote sensing of Australian rangelands. Remote Sens. Environ. 23:313-331. Graves, T. and J. Waller. 2006. Understanding the causes of missed global positioning system telemetry fixes. J. Wildl. Manage. 70:844-851. Green, W. C. H. 1992. Social influences on contact maintenance interactions of bison mothers and calves: group size and nearest neighbor distance. Anim. Behav. 43:775-785. 73 Harris, N. R., D. E. Johnson, N. K. McDougald, and George, M. R. 2007. Social associations and dominance of individuals in small herds of cattle. Rangeland Ecol. Manage. 60:339-349. Hart, R. H., J. Bissio, M. J. Samuel, and J. W. Waggoner. 1993. Grazing systems, pasture size, and cattle grazing behavior, distribution and gains. J. Range Manage. 46:81-87. Hart, R. H. and C. S. Hoveland. 1989. Objectives of grazing trials. Pages 1-5 In Grazing Research: Design, Methodology, and Analysis. G.C. Marten (ed) CSSA Spec. Pub. No. 16. Crop Sci. Soc. Amer., Madison, WI. Hart, R. H., M. J. Samuel, J. W. Waggoner Jr., and M. A. Smith. 1991. Grazing systems in Wyoming – Impacts of grazing pressure and livestock distribution. Rangelands 13:12-16. Harvey, T. L. and J. L. Launchbaugh. 1982. Effect of horn flies on behavior of cattle. J. Econ. Entomol. 75:25-27. Haydon, D. T., J. M. Morales, A. Yott, D.A. Jenkins, R. Rosatte, and J.M. Fryxell. 2008. Socially informed random walks: incorporating group dynamics into models of population spread and growth. Proc. Biol. Sci. 275:1101-1109. Havstad, K. M., E. L. Fredrickson, and R. E. Estell. 2004. Animal unit equivalents: an examination of the sheep to cattle ratio for stocking rangelands. http://usdaars.nmsu.edu/biblio/pdf/426.pdf. Accessed Oct. 2, 2009. Heard, D.C., L.M. Ciarniello, and D.R. Seip. 2008. Grizzly bear behavior and global positioning system collar fix rate. J. Wildl. Manage. 72:596-602. Henderson, C. R. 1959. Design and analysis of animal husbandry experiments. Pages 1-55 In Techniques and Procedures in Animal Production Research. C. E. Terrill (chair) Amer. Soc. Anim. Prod. Beltsville, MD. Hejcmanová, P., M. Stejskalová, V. Pavlu, and M. Hejcman. 2009. Behavioural patterns of heifers under intensive and extensive continuous grazing on speciesrich pasture in the Czech Republic. Appl. Anim. Behav. Sci. 117:137-143. Hemson, G., P. Johnson, A. South, R. Kenward, R. Ripley, and D. McDonald. 2005. Are kernels the mustard? Data from global positioning system (GPS) collars suggests problems for kernel home-range analyses with leastsquares cross-validation. J Anim. Ecol. 74:455-463. Hendrie, C. A. and S. Bennett. 1984. A microcomputer technique for the detailed behavioural, and automatic statistical, analysis of animal behaviour. Phys. Behav. 32:865-869. Hengl, T., E. E. van Loom, J. Shamoun-Baranes, and W. Bouten. 2008. Geostatistical analysis of GPS trajectory data: Space-time densities. Pages 17-24 In Proc. 8th Int. Symp. Spat. Acc. Asst. Nat. Res. Env. Sci. June 25-27. Shanghai, China. Hernandez, L., H. Barral, G. Halffter, and S. S. Colon. 1999. A note on the behavior of feral cattle in the Chihuahuan Desert of Mexico. Appl. Anim. Behav. Sci. 63:259-267. Herring, T. A. 1996. The global positioning system. Sci. Amer. 274:44-50. Hessle, A. K. 2009. Effects of social learning of foraging behaviour and live weight gain in first-season grazing calves. Appl. Anim. Behav. Sci. 116:150-155. Hill, M. J. 2004. Grazing agriculture – managed pastures, grassland and rangeland. Pages 449-530 In: Manual of Remote Sensing, Remote Sensing for Natural Resource Management and Environmental Monitoring Vol 4. S.L. Ustin (ed) John Wiley & Sons, Hoboken, NJ. Hinch, G. N., C. J. Thwaites, J. J. Lynch, and A .J. Pearson. 1982. Spatial relationships within a herd of young sterile bulls and steers. Appl. Anim. Ethol. 8:27-44. Hiroaki, M. and N. Takaaki. 2007. Grazing behavior of cows measured by handheld GPS and bite counter collar: A case of Fazenda Baía Bonita in South Pantanal, Brazil. People‟s Cultural Geography (translation). 59:30-42. Hobson, K. A. and L. I. Wassenaar (eds). 2008. Tracking Animal Migration with Stable Isotopes. Academic Press. San Diego, CA. Hodgson, R. E. 1933. Influence of pasture management upon the grazing habits of cattle. J. Agric. Res. 47:417424. Hooper, J. F., J. P. Workman, J. B. Grumbles, and C. W. Cook. 1969. Improved livestock distribution with fertilizer – a preliminary economic evaluation. J. Range Manage. 22:108-110. Horne, J. S. and E. O. Garton. 2006. Selecting the best home range model: An information-theoretic approach. Ecol. 87:1146-1152. Horne, J. S., E. O. Garton, and J. L. Rachlow. 2008. A synoptic model of animal space use: Simultaneous estimation of home range, habitat selection, and inter/intra-specific relationships. Ecol. Model. 214:338348. Howery, L. D., F. D. Provenza, R. E. Banner, and C. B. Scott. 1998. Social and environmental factors influence cattle distribution on rangeland. Appl. Anim. Behav. Sci. 55:231-244. Howery, L. D., D. W. Bailey, G. B. Ruyle, and W. J. Renken. 2000. Cattle use visual cues to track food locations. Appl. Anim. Behav. Sci. 67:1-14. Huber, R., R. Baumung, M. Wurzinger, D. Semambo, O. Mwai, and C. Winckler. 2008. Grazing, social and comfort behaviour of Ankole and crossbred (Ankole x Holstein) heifers on pasture in south western Uganda. Appl. Anim. Behav. Sci. 112:223-234. Hunt, L. P., S. Petty, R. Cowley, A. Fisher, A. J. Ash, and N. MacDonald. 2007. Factors affecting the management of cattle grazing distribution in northern Australia: Preliminary observations on the effect of paddock size and water points. Rangeland J. 29:169-179. Hurlbert, S. H. 1984. Pseudoreplication and the design of ecological field experiments. Ecol. Mono. 54:187-211. Hurn, J. 1993. Differential GPS explained. Trimble Navigation. Pt. No. 23036. Iliffe, J. 1999. Datums and Map Projections for Remote Sensing, GIS and Surveying. Whittles Publishing, Dubeath, Scotland. Immelmann, K. and C. Beer. 1989. A dictionary of ethology. Harvard Univ. Press, Cambridge, MA. 74 Kernohan, B. J., R. A. Gitzen, and J. J. Millspaugh. 2001. Analysis of animal space use and movements. Pages 125–166 In Radio Tracking and Animal Populations. J. J. Millspaugh and J. M. Marzluff (eds) Academic Press, San Diego, CA. Kie, J. G., J. A. Baldwin, and C. J. Evans.1996. CALHOME: a program for estimating animal home ranges. Wildl. Soc. Bull. 24:342-344. Kie, J .G. and B. B. Boroski. 1996. Cattle distribution, habitats, and diets in the Sierra Nevada of California. J. Range Manage. 49:482-488. Kiker, G. A., D. J. Clark, C. J. Martinez, and R. E. Shulze. 2006. A Java-based object oriented modeling system for Southern African hydrology: Verification in the Mgeni River Watershed, Kwazulu-Natal, South Africa. Trans. ASABE 49:1419-1433. Kinetic Resource and Environmental Spatial Systems modeler (KRESS). 2006. The KRESS Modeler Ver 2.0.1 User‟s Manual. http://kress.us/program.htm. Accessed Jan. 11, 2010. Krige, D. G. 1951. A statistical approach to some basic mine valuation problems on the Witwatersrand. J. Chem. Metal. Mining Soc. S. Afr. 52:119-139. Krysl, L. J., M. E. Branine, A. U. Cheema, M. A. Funk, and M. L. Galyean. 1989. Influence of soybean meal and sorghum grain supplementation on intake, digesta kinetics, ruminal fermentation, site and extent of digestion and microbial protein synthesis in beef steers grazing blue grama rangeland. J. Anim. Sci. 67:30403051. Ksiksi, T. and E. A. Laca. 2000. Can social interactions affect food searching efficiency of cattle? Rangeland J. 22:235-242. Kushwaha, S. P. S. and P. S. Roy. 2002. Geospatial technology for wildlife habitat evaluation. Trop. Ecol. 43:137-150. Laca, E. A. 2009. New approaches and tools for grazing management. Rangeland Ecol. Manage. 62:407-417. Landsberg, J., and J. Stol. 1996. Spatial distribution of sheep, feral goats and kangaroos in woody rangeland paddocks. Rangeland J. 18:270-291. Larson, M. A. 2001. A catalog of software to analyze radiotelemetry data. Pages 397-421 In Radio Tracking And Animal Populations. J. J. Millspaugh and J. M. Marzluff (eds) Academic Press, San Diego, CA. Lathrop, W. J., D. D. Kress, K .M. Havstad, D. E. Doornbos, and E. L. Ayers. 1988. Grazing behavior of rangeland beef cows differing in milk production. Appl. Anim. Behav. Sci. 315-327. Lazo, A. 1994. Social segregation and the maintenance of social stability in a feral cattle population. Anim. Behav. 48:1133-1141. Lazo, A. and R. C. Soriguer. 1993. Size-biased foraging behaviour in feral cattle. Appl. Anim. Behav. Sci. 36:99110. Lee, J. and D. W. S. Wong. 2001. Statistical Analysis With ArcView GIS. John Wiley and Sons, Inc. Hoboken, NJ. Le Munyan, C. D., W. White, E. Nybert, and J. J. Christian. 1959. Design of a miniature radio transmitter for use in animal studies. J. Wildl. Manage. 23:107-110. Jackson, R. D. 1986. Remote sensing of biotic and abiotic plant stress. Ann. Rev. Phytopathol. 24:265-287. James, C. D., J. Laandsberg, and S. R. Morton. 1999. Provision of watering points in the Australian arid zone: a review of effects on biota. J. Arid Environ. 41:87-121. Jeung, H., Q. Liu, H. T. Shen, and X. Zhou. 2008. A hybrid prediction model for moving objects. Pages 70-79 in IEEE 24th International Conference on Data Engineering (ICDE). Mexico, Cancún. Johnson, A. R., J. A. Wiens, B. T. Milne, and T. O. Crist. 1992. Animal movements and population dynamics in heterogeneous landscapes. Landscape Ecol. 7:63-75. Johnson, D. D. P., R. Kays, P. G. Blackwell, and D.W. Macdonald. 2002. Does the resource dispersion hypothesis explain group living? Trends Ecol. Evol. 17:563-570. Johnson, L. B. 1990. Analyzing spatial and temporal phenomena using geographical information systems. Landscape Ecol. 4:31-43. Jung, J. N. 1994. Applied System Identification. Prentice Hall. Englewood Cliffs, NJ. Kao, S. D., F. Z. Shaw, M. S. Young and G. J. Jan. 1995. A new automated method for detection and recording of animal moving path. J. Neurosci. Meth. 63:205-209. Kareiva, P. and N. Shigesada. 1983. Analysing insect movement as a correlated random walk. Oecologia 56:234-238. Karnieli, A., U. Gilad, M. Ponzet, T. Svoray, R. Mirzadinov, and O. Fedorina. 2008. Assessing landcover change and degradation in the Central Asian deserts using satellite image processing and geostatistical methods. J. Arid Environ. 72:2093-2105. Kawamura, K., T. Akiyama, H. Yokota, M. Tsutsumi, T. Yasuda, O. Watanabe, and S. Wang. 2005. Quantifying grazing intensities using geographic information systems and satellite remote sensing in the Xilingool steppe region, Inner Mongolia, China. Agri. Ecosyst. Environ. 107:83-93. Keating, K. A., W. G. Brewster, and C. H. Key. 1991. Telemetry: Performance of animal-tracking systems. J. Wildl. Manage. 55:160-171. Keister, G. P. Jr., C. E. Trainer, and M. J. Willis. 1988. A self-adjusting collar for young ungulates. Wildl. Soc. Bull. 16:321-323. Kellerman, A. 1981. Centrographic measures in geography: Concepts and techniques in modern Geography (CATMOG) No. 32. Norwich: Geo Books, Univ. East Angola. Kennedy, M. and R. D. Gray. 1993. Can ecological theory predict the distribution of foraging animals? A critical analysis of experiments on the Ideal Free Distribution. Oikos. 68:158-166. Kenward, R. 1987. Wildlife Radio Tagging: Equipment, Field Techniques and Data Analysis (Biological Techniques). Academic Press. London, UK. Kenward, R. E. 2001a. A Manual For Wildlife Radio Tagging. Academic Press, San Diego, CA. Kenward, R. E. 2001b. Historical and practical perspectives. Page 3-12 In Radio Tracking And Animal Populations. J. J. Millspaugh and J. M. Marzluff (eds) Academic Press, San Diego, CA. 75 Lowe, J. C., S. M. Abeyesinghe, T. G. M. Demmers, C. M. Wathes, and D. E. F. McKeegan. 2007. A novel telemetric logging system for recording physiological signals in unrestrained animals. Comp. Elect. Agric. 57:74-79. Ludwig, J. A., G. N. Bastin, R. W. Eager, R. Karfs, P. Ketner, and G. Pearce. 2000. Monitoring Australian rangeland sites using landscape function indicators and ground-and remote-based techniques. Environ. Monitoring Assess. 64:167-178. Ludwig, J. A., D. J. Tongway, G. N. Bastin, and C. D. James. 2004. Monitoring ecological indicators of rangeland functional integrity and their relation to biodiversity at local to regional scales. Austral Ecol. 29:108-120. Macdonald, D. W. and D. D. P. Johnson. 2001. Dispersal in theory and practice: Consequences for conservation biology. Pages 358-372 In Dispersal. J. Clobert, E. Danchin, A. A. Dhondt, and J. D. Nichols (eds) Oxford Univ. Press, Oxford, UK. Malechek, J. C. and B. M. Smith. 1976. Behavior of range cows in response to winter weather. J. Range Manage. 29:9-12. Manly, B. F., L. L. McDonald, and D.L. Thomas. 1993. Resource selection by animals: statistical design and analysis for field studies. Chapman and Hall, London, UK. Mardia, K. V. 1972. Statistics Of Directional Data. Academic Press. London, UK. Marion, G., D. L. Swain, and M. R. Hutchings. 2005. Understanding foraging behaviour in spatially heterogeneous environments. J. Theor. Biol. 232:127142. Marsh, L. M. and R. E. Jones. 1988. The form and consequences of random walk movement models. J. Theor. Biol. 133:113-131. Martin, S. C. and D. E. Ward. 1973. Salt and meal-salt help distribute cattle use on semidesert range. J. Range Manage. 26:94-97. Marzluff, J. M., S. T. Knick, and J. J. Millspaugh. 2001. Hightech behavioral ecology: modeling the distribution of animal activities to better understand wildlife space use and resource selection. Pages 309–326 In Radio Tracking And Animal Populations. J. J. Millspaugh and J. M. Marzluff (eds) Academic Press, San Diego, CA. Matheron, G. 1965. Les Variables Régionalisées et leur Estimation. Masson, Paris. Matthiopoulos, J. 2003. The use of space by animals as a function of accessibility and preference. Ecol. Model. 159:239-268. McIlvain, E. H. and M. C. Shoop. 1965. Methods for distributing livestock grazing pressure. Progress in range research in 17 western states. H. J. Hodgson compiler. Coop. State Res. Ser. USDA, Washington DC. Millspaugh, J. J. and Marzluff, J. M. (eds). 2001. Radio Tracking and Animal Populations. Academic Press. San Diego, CA. Milne, B. T. 1992. Spatial aggregation and neutral models in fractal landscapes. Amer. Nat. 139:32-57. Lehner, P. N. 1979. Handbook Of Ethological Methods. Garland STPM Press, New York, NY. Leyequien, E., J. Verrelst, M. Slot, G. Schaepman-Sstrub, I. M. A. Heitkönig, and A. Skidmore. 2007. Capturing the fugitive: Applying remote sensing to terrestrial animal distribution and diversity. Int. J. Appl. Earth Observ. Geoinform. 9:1-20. Li, Z. and W. Wang. 1998a. Dispersal models of describing grassland resource and the radial type pattern of grazing animals I. J. Biomath. 13:1102-1107. Li, Z. and W. Wang. 1998b. Dispersal models of describing grassland resource and the radial type pattern of grazing animals II. J. Biomath. 13:1108-1115. Li, Z. and W. Wang. 2006. Distribution models of grazing animals around a grassland resource point. Appl. Math. Comp. 172:925-937. Linnane, M. I., A. J. Brereton, and P. S. Giller. 2001. Seasonal changes in circadian grazing patterns of Kerry cows (Bos taurus) in semi-feral conditions in Killarney National Park, Co. Kerry Ireland. Appl. Anim. Behav. Sci. 71:277-292. Littell, R. C., G. A. Milliken, W. W. Stroup, and R. D. Wolfinger. 1996. SAS System for Mixed Models. SAS Institute Inc., Cary, NC. Ljung, L. 1999. System Identification: Theory for the User. Prentice-Hall. Upper Saddle River, NJ. Loehle, C. 1990. Home range: A fractal approach. Landscape Ecol. 5:39-52. Longley, P. A., M. F. Goodchild, D. J. Maguire, and D. W. Rhind. 2005. Geographic Information Systems: Principles, Techniques, Management and Applications. Abridged edition. J. Wiley and Sons. Hoboken, NJ. Loeb, J. 1918. Forced Movements, Tropisms and Animal Conduct. J .B. Lippincott Co. Philadelphia, PA. Loewer, O. J. 1989. Issues on modeling grazing systems. Pages 127-136 In Grazing Research: Design, Methodology, and Analysis. G.C. Marten (ed) CSSA Special Publ. No. 16. Crop Sci. Soc. Amer. Madison, WI. Lock, R., I. Rademacher, H. Nonn, and W. Kühbauch. 2004. Methods of digital image processing to quantify ground cover of turf grass. Pages 790-792 In Grassland Science Eur. Vol 9. A. Lüscher, B. Jeangros, W. Kessler, O. Huguenin, M. Lobsinger, N. Milllar, and D. Suter (eds) Eur. Grassland Fed. Zurich, Switzerland. Long, E. S., D. R. Diefenbach, C. S. Rosenberry, B. D. Wallingford, and M. D. Grund. 2005. Forest cover influences dispersal distance of white-tailed deer. J. Mamm. 86:623-629. Low, W.A. (ed). 1978. The physical and biological features of Kunoth Paddock in central Australia. CSIRO (Australia) Div. Land Resources Manage. Tech. Papers. Canberra, Austral. Low, W. A., M. L. Dudzinski, and W. J. Müller. 1981. The influence of forage and climatic conditions on range community preference of Shorthorn cattle in arid central Australia. J. Appl. Ecol. 18:11-26. 76 Okubo, A. 1980. Diffusion and ecological problems: Mathematical models. Springer-Verlag. New York, NY. O‟Shea, J. 1969. Evaluation of a simple device for measuring the time animals spend grazing. Irish J. Agri. Res. 8:329-335. Otis, D. L. and G. C. White. 1999. Autocorrelation of location estimates and the analysis of radiotracking data. J. Wildl. Manage. 63:1039-1044. Palancioğlu, H. M. 2003. Extracting movement patterns using fuzzy and neuron-fuzzy approaches. Ph.D. Diss. Univ. of Maine. Orono. Pandey, V. 2007. Analysis and modeling of cattle distribution in complex agro-ecosystems of south Florida. Ph.D. Diss. Univ. of Florida, Gainesville. Paraan, N. C. and J. P. Esguerra. 2006. Exact moments in a continuous time random walk with complete memory of it history. Phys. Rev. E74:032101-1-032101-4. Parunak, H. V. D., S. A. Brueckner, R. Matthews, and J. Sauter. 2006. Swarming methods for geospatial reasoning. Int. J. Geog. Info. Sci. 20:945-964. Parr, A. E. 1927. A contribution to the theoretical analysis of the schooling behavior of fishes. Occasional Pap. Bingham Oceanography Coll. 1:1-32. Pedersen, S. and C. B. Pedersen. 1995. Animal activity measured by infrared detectors. J. Agric. Eng. Res. 61:239-246. Peterson, R. A. and E. J. Woolfolk. 1955. Behavior of Hereford cows and calves on short grass range. J. Range Manage. 8:51-57. Petersen, R. G. and H. L. Lucas. 1960. Experimental errors in grazing trials. Pages 747-750 In Proc. 8th Int. Grassland Congr. Reading, England. July 11-21. Alden Press, Oxford, UK. Petrusevics, P. E. and Davisson, T. W. 1975. A biotelemetry system for cattle behaviour studies. Proc. IREE. 36:72-76. Phillips, R., O. Beeri, E. Scholljegerdes, D. Bjergaard, and J. Hendrickson. 2009. Integration of geospatial and cattle nutrition information to estimate paddock grazing capacity in Northern US prairie. Agric. Syst. 100:72-79. Pickup, G. 1994. Modeling patterns of defoliation by grazing animals in rangelands. J. Appl. Ecol. 31:231246. Pickup, G. and G. N. Bastin. 1997. Spatial distribution of cattle in arid rangelands as detected by patterns of change in vegetation cover. J. Appl. Ecol. 34:657-667. Pickup, G. and V. H. Chewings. 1988. Estimating the distribution of grazing and patterns of cattle movement in a large arid zone paddock. An approach using animal distribution models and Landsat imagery. Int. J. Remote Sens. 9:1469-1490. Pickup, G., G. N. Bastin, and V. H. Chewings. 1994. Remote-sensing-based condition assessment for nonequilibrium rangelands under large-scale commercial grazing. Ecol. Appl. 4:497-517. Pinchak, W. E., M. A. Smith, R. H. Hart, and J. W. Waggoner, Jr. 1991. Beef cattle distribution patterns on foothill range. J. Range Manage. 44:267-275. Pinheiro, J. C. and D. M. Bates. 2000. Theory and computational methods for LME models. Pages 56-96 In Mixed Effects Model in S and S-Plus. Springer, NY. Mirabet, V., P. Auger, and C. Lett. 2007. Spatial structures in simulations of animal grouping. Ecol Model. 201:468-476. Mitchell, J. E. and R. T. Rogers. 1985. Food habits and distribution of cattle on a forest and pasture range in Northern Idaho. J. Range Manage. 38:214-220. Mohr, C. O. 1947. Table of equivalent populations of North American small mammals. Amer. Midland Natur. 37:223-249. Moll, R. J., J. J. Millspaugh, J. Beringer, J. Sartwell, Z. He, J.A. Eggert, and X. Zhao. 2009. A terrestrial animalborne video system for large mammals. Comp. Elect. Agric. 66:133-139. Moorcroft, P. R. and M. A. Lewis. 2006. Mechanistic home range analysis. Monograph in Population Biology 43. Princeton Univ. Press. Princeton, NJ. Moore, R. J. and R. T. Clarke. 1983. A distribution function approached to modeling basin sediment yield. J. Hydrol. 65:239-257. Morales, J. and S. Ellner. 2002. Scaling up animal movements in heterogeneous landscapes: The importance of behavior. Ecol. 83:2240-2247. Moreau, M., S. Siebert, A. Buerkert, and E. Schlecht. 2009. Use of a tri-axial accelerometer for automated recording and classification of goats‟ grazing behaviour. Appl. Anim. Behav. Sci. 119:158-170. Morley, F. H. W. (ed). 1981. Grazing Animals. Elsevier. New York, NY. Müller, R. and L. Schrader. 2003. A new method to measure behavioural activity levels in dairy cows. Appl. Anim. Behav. Sci. 83:247-258. Murguia, M. and J. L. Villasenor. 2000. Estimating the quality of the records used in quantitative biogeography with presence-absence matrices. Ann. Bot. Fennici. 37:289-296. Murphy, K. 1998. Hidden Markov Model (HMM) Toolbox. http://people.cs.ubc.ca/~murphyk/Software/HMM/hmm. html. Accessed Dec. 08, 2009. Nadimi, E. S. and H. T. Søgaard. 2009. Observer Kalman filter identification and multiple-model adaptive estimation technique for classifying animal behaviour using wireless sensor networks. Comp. Elect. Agric. 68:9-17. Naito, Y. 2004. Bio-logging science. Mem. Natl. Inst. Polar Res. Spec. Issue 58:118-132. Neft, D. S. 1966. Statistical analysis for areal distributions. Monograph Series No. 2. Reg. Sci. Res. Inst. Philadelphia, PA. Nerini, D., P. Monestiez, and C. Mant'e. 2010. Cokriging for spatial functional data. J. Multivariate Anal. 101:409-418. Nielsen, B. L. 1999. On the interpretation of feeding behaviour measures and the use of feeding rate as an indicator of social constraint. Appl. Anim. Behav. Sci. 63:79-91. Norton, B. E. and P. S. Johnson. 1986. Impact of grazing on crested wheatgrass in relation to plant size. Pages 275279 In Crested Wheatgrass: Irs Values, Problems, and Myths. K. Johnson (ed) Symp. Proc. Utah State Univ. Logan, UT. 77 Pollak, E. R., D. W. Bailey, and R. Weinmeister. 2008. Social interactions of hill-climber and bottom-dweller cows in small and large herds. Abstr. Soc. Range Manage. Ann. Meeting, January 27-31, Louisville, KY. Soc. Range Manage. Denver, CO. Porath, M. L., P. A. Momont, T. DelCurto, N. R. Rimbey, J. A. Tanaka, and M. McInnis. 2002. Offstream water and trace mineral salt as management strategies for improved cattle distribution. J. Anim. Sci. 80:346-356. Powell, J. and T. W. Box. 1966. Brush management influences preference values of south Texas woody species for deer and cattle. J. Range Manage. 19:212214. Price, M. 2006. Mastering ArcGIS. 2nd Ed. McGraw-Hill, New York, NY. Pringle, H. J. R. and J. Landsberg. 2004. Predicting the distribution of livestock grazing pressure in rangelands. Austral. Ecol. 29:31-39. Puers, B., P. Wouters, and M. DeCooman. 1993. Low power multi-channel sensor interface for use in digital telemetry. Sens. Actu. A Physical 37-38:260-267. R Development Core Team. 2006. R: A language and environment for statistical computing. R Foundation for statistical computing. Vienna, Austria. http://www.rproject.org/ Accessed Dec. 01, 2009. Rahman, A. F., J. A. Gamon, D. A. Sims, and M. Schmidts. 2003. Optimum pixel size for hyperspectral studies of ecosystem function in southern California chaparral and grassland. Remote Sens. Environ. 84:192-207. Ramsey, F. L. and K. Harrison. 2004. A closer look at detectability. Environ. Ecol. Stat. 11:73-84. Rands, S. A., G. Cowlishaw, R. A. Pettifor, J. M. Rowcliffe, and R .A. Johnstone. 2008. The emergence of leaders and followers in foraging pairs when the qualities of individuals differ. BMC Evol. Biol. 8:51-68. Read, J. M., C. P. D. Birch, and J. A. Milne. 2002. HeathMod: A model of the impact of seasonal grazing by sheep on upland heaths dominated by Calluna volgaris (heather). Biol. Cons. 105:279-292. Reid, D. 1951. A quantitative method for determining palatability of pasture plants. J. Brit. Grassland Soc. 6:187-195. Reiss, M. 1988. Scaling of home range size: Body size, metabolic needs and ecology. Trends Ecol. Evol. 3:8586. Reinhardt, V. and A. Reinhardt. 1981. Cohesive relationships in a cattle herd (Bos indicus). Behav. 77:121-151. Reluga, T. C. and S. Viscido. 2005. Simulated evolution of selfish herd behavior. J. Theor. Biol. 234:213-225. Rempel, R. S., A. R. Rodgers, and K. F. Abraham. 1995. Performance of a GPS animal location system under boreal forest canopy. J. Wildl. Manage. 59:543-551. Rempel, R. S. and A. R. Rogers. 1997. Effects of differential correction on accuracy of a GPS animal location system. J. Wildl. Manage. 61:525-530. Renken, W. J., L. D. Howery, G. B. Ruyle, and R .M. Enns. 2008. Cattle generalize visual cues from the pen to the field to select initial feeding patches. Appl. Anim. Behav. Sci. 109:128-140. Riewe, M.E. 1961. Use of the relationship of stocking rate to gain of cattle in an experimental design for grazing trials. Agron. J. 53:309-313. Renolen, A. 1999. Concepts and methods for modeling temporal and spatiotemporal information. Ph.D. Diss. Norwegian Univ. of Sci. and Techol. (NTNU), Trondheim. Roath, L. R. and W. C. Krueger. 1982. Cattle grazing influence on a mountain riparian zone. J. Range Manage. 35:100-103. Robert, B., B. J. White, D. G. Renter, and R. L. Larson. 2009. Evaluation of three-dimensional accelerometers to monitor and classify behavior patterns in cattle. Comp. Elect. Agric. 67:80-84. Roberts, C. M. 2006. Radio frequency identification (RFID). Comp. Sec. 25:18-26. Röder, A., T. Kuemmerle, J. Hill, V.P. Papanastasis, and G. M. Tsourlis. 2007. Adaptation of a grazing gradient concept to heterogeneous Mediterranean rangelands using cost surface modeling. Ecol. Model. 204:387-398. Rogers, A. R. 2001. Recent telemetry technology. Pages 79-121 In: Radio Tracking And Animal Populations. J. J. Millspaugh and J. M. Marzluff (eds) Academic Press, San Diego, CA. Rogers, A. R. and Carr, A. P. 2002. HRE: The home range extension for ArcView. User‟s manual. Centre for Northern Forest Ecosystem Research. Ontario Ministry of Natural Resources. Canada. Roloff, G. J., M. L. Donovan, D. W. Linden, and M. L. Strong. 2009. Lessons learned from using GIS to model landscape-level wildlife habitat. Pages 287-320 In Models for Planning Wildlife Conservation in Large Landscapes. J. J. Millspaugh and F. R. Thompson, III (eds) Academic Press. Burlington, MA. Rook, A. J., B. Dumont, J. Isselstein, K. Osoro, M. F. Wallis, W. DeVries, G. Parente, and J. Mills. 2004. Matching type of livestock to desired biodiversity outcomes in pastures – a review. Biol. Cons. 119:137150. Ropert-Coudert, Y. and R. P. Wilson. 2005. Trends and perspectives in animal-attached remote sensing. Front. Ecol. Environ. 3:437-444. Roshier, D. A. and H. I. Nicol. 1998. Implications of spatiotemporal variation in forage production and utilization for animal productivity in extensive grazing systems. Rangeland J. 20:3-25. Rutter, M. 2007. The integration of GPS, vegetation mapping and GIS in ecological and behavioural studies. R. Bras. Zootec. 36(suppl):63-70. Sampson, B. A. and G. D. Delgiudice. 2006. Tracking the rapid pace of GIS-related capabilities and their accessibility. Wildl. Soc. Bull. 34:1446-1454. SAS Institute. 1993. SAS Technical Report P-243: The Genmod Procedure. SAS Institute Inc. Cary, NC. SAS Institute. 2000. SAS/STAT User‟s Guide, Version 8, Vol. II. SAS Institute Inc. Cary, NC. SAS Institute. 2002. JMP Statistics and Graphics Guide, Version 5. SAS Institute Inc. Cary, NC. Samuel, M. J., F. Rauzi, and R. H. Hart. 1980. Nitrogen fertilization of range: yield, protein content, and cattle behavior. J. Range Manage. 33:119-121. 78 Senft, R. L. 1989. Hierarchial foraging models: Effects of stocking and landscape composition on simulated resource use by cattle. Ecol. Model. 46:2283-303. Senft, R. L., M. B. Coughenour, D. W. Bailey, L. R. Rittenhouse, O. E. Shala, and D. W. Swift. 1987. Large herbivore foraging and ecological hierarchies. BioSci. 37:789-799. Senft, R. L., L. R. Rittenhouse, and R. G. Woodmansee. 1983. The use of regression models to predict spatial patterns of cattle behavior. J. Range Manage. 36:553557. Senft, R. L., L. R. Rittenhouse, and R. G. Woodmansee. 1985. Factors influencing patterns of cattle grazing behavior on shortgrass steppe. J. Range Manage. 38:8287. Seton. E. T. 1909. Life-histories of northern animals. An account of the mammals of Manitoba. Charles Scribner‟s Sons. New York, NY. Sevi, A., A. Muscio, D. Dantone, V. Iascone, and F. D‟Emilio. 2001. Paddock shape effects on grazing behavior and efficiency in sheep. J Range Manage. 54:122-125. Shannon, C.E. and W. Weaver. 1949. The mathematical theory of communication. Univ. of Illinois. Urbana. Shaw, D. M. and S. F. Atkinson. 1990. An introduction to the use of geographic information systems for ornithological research. The Condor. 92:564-570. Sheehy, C. M. 2007. Comparison of spatial distribution and resource use by Spanish and British breed cattle in Northeastern Oregon Prairie ecosystems. MS Thesis. Oregon State Univ. Corvallis. Sherwin, C. 1994. Farm Animals. Pages 125-144 in Video Techniques in Animal Ecology and Behaviour. S.D. Wratten (ed) Chapman and Hill, London, UK. Shiyomi, M. 2004. How are distances between individuals of grazing cows explained by a statistical model? Ecol. Model. 172:87-94. Shiyomi, M. and M. Tsuiki. 1999. Model for the spatial pattern formed by a small herd in grazing cattle. Ecol. Model. 119:231-238. Sinclair, A. R .E. 1992. Do large mammals disperse like small mammals? Pages 229-239 In Animal Dispersal: Small Mammals as a Model. N. C. Stenseth and W. Z. Lidicker, Jr., (eds) Chapman & Hall, London, UK. Skovlin, J. M. 1957. Range riding – the key to range management. J. Range Manage. 10:269-271. Smith, D. R. and R. L. Lang. 1958. The effect of nitrogenous fertilizers on cattle distribution on mountain range. J. Range Manage. 11:218-249. Snedecor, G. W. and W .G. Cochran. 1967. Statistical Methods. 6th ed. Iowa State Univ. Press. Ames, IA. Sørensen, J. T. 1990. Validation of livestock herd simulation models: A review. Livst. Prod. Sci. 26:79-90. Sowell, B. F., J. C. Mosley, and J.G.P. Bowman. 2000. Social behavior of grazing beef cattle: Implications for management. J. Anim. Sci. 77:1-6. Springer, J. T. 1979. Some sources of bias and sampling error in radio triangulation. J. Wildl. Manage. 43:926935. Sato, S., I. Ito, and K. Hayashi. 1976. Studies on the spatial pattern of grazing cattle. J. Jap. Grassland Sci. 22:313318. Schacht, W. H., A. Guru, P. E. Reece, J. D. Volesky, and D. Cotton. 2005. Using geospatial information technologies and field research to enhance classroom learning. J. Nat. Res. Life Sci. Educ. 34:62-66. Schauer, C. S., D. W. Bohnert, D. C. Ganskopp, C. J. Richards, and S. J. Falck. 2005. Influence of protein supplementation frequency on cows consuming lowquality forage: Performance, grazing behavior, and variation in supplement intake. J. Anim. Sci. 83:17151725. Schellinck, J. A. 2008. A general perception based framework for modeling animal aggregation. Ph.D. Diss. Carletion Univ., Ottawa, Ontario, Canada. Schlecht, E., C. Hülsebusch, F. Mahler, and K. Becker. 2004. The use of differentially corrected global positioning system to monitor activities of cattle at pasture. Appl. Anim. Behav. Sci. 85:185-202. Schlecht, E., P. Hiernaux, I. Kadaouré, C. Hülsebusch, and F. Mahler. 2006. A spatio-temporal analysis of forage availability and grazing and excretion behaviour of herded and free grazing cattle, sheep and goats in Western Niger. Agric. Ecosys. Environ. 113:226-242. Schellberg, J., M. J. Hill, R. Gerhards, M. Rothmund and M. Braun. 2008. Precision agriculture on grassland: Applications, perspectives and constraints. Euro. J. Agron. 29:59-71. Schmidt, P. J. 1969. Observations on the behaviour of cattle in a hot dry region of the Northern Territory of Australia, with particular reference to walking, watering and grazing. M.Ru.Sc. Thesis. The University of New England, Armidale. Schwager, M., D. M. Anderson, Z. Butler, and D. Rus. 2007. Robust classification of animal tracking data. Comp. Electr. Agric. 56:46-59. Schwager, M., C. Detweiler, I. Vasilescu, D.M. Anderson, and D. Rus. 2008. Data-driven identification of group dynamics for motion prediction and control. J. Field Robot. 25:305-324. Scifres, C. J., J. R. Scifres, and M. M. Kothmann. 1983. Differential grazing use of herbicide-treated areas by cattle. J. Range Manage. 36:65-69. Seaman, D. E. and R. A. Powell. 1996. An evaluation of the accuracy of kernel density estimators for home range analysis. Ecol. 77:2075-2085. Seaman, D. E., B. Griffith, and R. A. Powel. 1998. KERNELHR: A program for estimating animal home ranges. Wildl. Soc. Bull. 26:95-100. Seaman, E. D., J. J. Millspaugh, B. J. Kernohan, G. C. Brundige, K. J. Raedeke, and R. A. Gitzen. 1999. Effects of sample size on kernel home range estimates. J. Wildl. Manage. 63:739-747. Searle, K. R. and L. A. Shipley. 2008. The comparative feeding behaviour of large browsing and grazing herbivores. Pages 117-148 In The Ecology of Browsing and Grazing. I.J. Gordon and H.H.T. Prins (eds) Springer-Verlag, Heidelberg, Germany. 79 Squires, V. R. 1975. Social behaviour in domestic livestock: the basis for improved animal husbandry. Appl. Anim. Ethol. 1:177-184. Squires, V. R. 1978. Effects of management and the environment on animal dispersion under free range conditions. Pages 323-334 In Proc. 1st Animal World Congress Ethol. Appl. Zootech. Garsi, Madrid, Spain. Stafford, J. V. 2000. Implementing precision agriculture in the 21st century. J. Agric. Eng. Res. 76:267-275. Stafford-Smith, M. 1988. Modeling: three approaches to predicting how herbivores impact is distributed in rangelands. NM Agric. Exp. Sta. Reg. Res. Rep. 628. Starks, P. J., S. W. Coleman, and W. A. Phillips. 2004. Determination of forage chemical composition using remote sensing. J. Range Manage. 57:635-640. Stein, M. L. 1999. Interpolation of Spatial Data: Some Theory for Kriging. Springer, New York, NY. Stewart, K. M., R. T. Bowyer, J. G. Kie, N. J. Cimon, and B. K. Johnson. 2002. Temporspatial distributions of elk, mule deer, and cattle: Resource partitioning and competitive displacement. J. Mammal. 83:244-244. Stenseth, N. C. and W. Z. Lidicker, Jr. (eds). 1992. Animal Dispersal: Small Mammals as a Model. Chapman & Hall, London, UK. Stobbs, T. H. 1970. Automatic measurement of grazing time by dairy cows on tropical grass and legume pastures. Trop. Grassland 4:237-344. Stockwell, D. R. B. and I. R. Noble. 1992. Induction of sets of rules from animal distribution data: A robust and informative method of data analysis. Math. Comp. Simul. 33:385-390. Stone, M. and R. J. Brooks. 1990. Continuum regression: Cross–validated sequentially constructed prediction embracing ordinary least squares, partial least squares and principal components regression (with Discussion). J. Royal Stat. Soc. B. 52:237-239. Stricklin, W. R. and C. C. Kautz-Scanavy. 1984. The role of behavior in cattle production: A review of research. Appl. Anim. Ethol. 11:359-390. Stricklin, W. R. and J. A. Mench. 1987. Social organization. Pages 307-322 In Vet. Clinics N. Amer. Food Anim. Pract. Vol. 3 No. 2. W.B. Saunders Co., Philadelphia, PA. Stricklin, W. R., J. Z. Zhou, and H. W. Gonyou. 1995. Selfish animats and robot ethology: using artificial animals to investigate social and spatial behavior. Appl. Anim. Behav. Sci. 44:187-203. Stuth, J. W. 1991. Foraging behavior. Pages 65-83 In Grazing Management and Ecological Perspectives. R. K. Heitschmidt and J.W. Stuth (eds) Timber Press Inc. Portland, OR. Sui, D. Z. 2004. Tobler‟s First Law of Geography: A big idea for a small world. Ann. Assoc. Amer. Geographers 94:269-277. Sumpter, D. J. T. 2006. The principles of collective animal behavior. Phil. Trans. R. Soc. B. 361:5-22. Swain, D.L., T. Wark, and G.J. Bishop-Hurley. 2008. Using high fix rate GPS data to determine the relationships between fix rate, prediction errors and patch selection. Ecol. Mod. 212:273-279. Swihart, R. K. and N. A. Slade. 1985. Testing for independence of observations in animal movements. Ecol. 66:1176-1184. SYSTAT. 1992. „SYSTAT: Statistics, Ver. 5.2 Ed. SYSTAT Inc. Evanston, IL. Tate, K. W., E. R. Atwill, N. K. McDougald, and M .R. George. 2003. Spatial and temporal patterns of cattle feces deposition on rangeland. J. Range Manage. 56:432-438. Thrash, I. and J.F. Derry. 1999. Review of literature on the nature and modeling of piospheres. Koedoe. 42:73-94. Tobin, J. 1958. Estimation of relationships for limited dependent variables. Econometrica 26:24-36. Tomkins, N. and P. O‟Reagain. 2007. Global positioning systems indicate landscape preferences of cattle in the subtropical savannas. Rangeland J. 29:217-222. Tomkins, N. W., P. J. O‟Reagain, D. Swain, G. BishopHurley, and E. Charmley. 2009. Determining the effect of stocking rate on the spatial distribution of cattle for the subtropical savannas. Rangeland J. 31:267-276. Trail, L. W. and R. C. Bigalke. 2006. A presence-only habitat suitability model for large grazing African ungulates and its utility for wildlife management. Afr. J. Ecol. 45:347-354. Trénel, P., M. B. Jensen, E. L. Decker, and F. Skjøth. 2009. Quantifying and characterizing behavior in dairy calves using the IceTag automatic recording device. J. Dairy Sci. 92:3397-3401. Tribe, D.E. 1980. Animal production in the 21 st century. J. Aust. Inst. Agric. Sci. 46:36-40. Trotter, M. 2009. GPS livestock tracking (UNETracker). 13th Ann. Symp. Prec. Agric. Austral. Univ. of New England Armidale, NSW, Australia. http://www.une.edu.au/parg/UNETracker.php Accessed Dec. 03, 2009. Turchin, P. 1996. Fractal analyses of animal movement: A critique. Ecol. 77:2086-2090. Tucker, C. S., A. R. Rogers, and K. E. Schütz. 2008. Effect of solar radiation on dairy cattle behaviour, use of shade and body temperature in a pasture-based system. Appl. Anim. Behav. Sci. 109:141-154. Turner, L. W., M. Anderson, B. T. Larson, and M. C. Udal. 2001. Global positioning systems (GPS) and grazing behavior in cattle. Pages 640-650, paper number 701P0201 In: Livestock Environment VI: Proc. 6th Internatl. Symp. R. R. Stowell, R. Bucklin, and R. W. Bottcher (eds) ASABE, Louisville, KY. Turner, M. D., P. Hiernaux, and E. Schlecht. 2005. The distribution of grazing pressure in relation to vegetation resources in semi-arid West Africa: The role of herding. Ecosyst. 8:668-681. Umstatter, C., A. Waterhouse, and J. Holland. 2006. An automated method of simple behavior classification as a tool for management improvement in extensive systems. Pages 57-58 In: Proc. 4th Internatl. Workshop on Smart Sensors in Livest. Monitoring. Gargnano, Italy. Underwood, A. J. and M. G. Chapman. 1985. Multifactorial analyses of directions of movement of animals. J. Exp. Mar. Biol. Ecol. 91:17-43. 80 Ungar, E. D., Z. Henkin, M. Gutman, A. Dolve, A. Genizi, D. Ganskopp. 2005. Inference of animal activity from GPS collar data on free-ranging cattle. Rangeland Ecol. Manage. 58:256-266. Unwin, D. J. 2009. Statistics, Spatial. Pages 452-457 In: International. Encyclopedia of Human Geography. R. Kitchin and N. Thrift (eds) Elsevier. Maryland Heights, MO. Van Dyne, G. M., N. R. Brockington, Z. Szocs, J. Duek, and C. A. Ribic. 1980. Large herbivore subsystem. Pages 269-537 In: Grasslands, Systems Analysis and Man. A.I. Breymeyer and G.M. Van Dyne (eds) Cambridge Univ. Press. New York, NY. Valentine, K. A. 1947. Distance from water as a factor in grazing capacity of rangeland. J. For. 45:749-754. VanWagoner, H. C., D. W. Bailey, D. D. Kress, D. C. Anderson, and K. C. Davis. 2006. Difference among beef sire breeds and relationships between terrain use and performance when daughters graze foothill rangelands as cows. Appl. Anim. Behav. Sci. 97:105121. Vavra, M. and D. Ganskopp. 1998. Grazing behavior in ungulates: Current concepts and future challenges. Ann. Arid Zone. 37:319-335. von Hünerbein, K., and E. Rüter. 2001. A very lightweight flight recorder for homing pigeons based on GPS. Page 95 In: Proc. Conf. „Tracking animals with GPS,‟ March 12-13. Aberdeen, Scotland. Vuilleumier, S. and R. Metzger. 2006. Animal dispersal modeling: handling landscape features and related animal choices. Ecol. Mod. 190:159-170. Wade, T. G., B. W. Schultz, J. D. Wickham, and D. F. Bradford. 1998. Modeling the potential spatial distribution of beef cattle grazing using a geographic Information System. J. Arid Environ. 38:325-334. Wade, T. G., J. D. Wickham, M. S. Nash, A. C. Neale, K. H. Riitters, and K. B. Jones. 2003. A comparison of vector and raster GIS methods for calculating landscape metrics used in environmental assessments. Photogrammetric Eng. Remote Sens. 69:1399-1405. Walburger, K. J., M. Wells, M. Vavra, T. Delcurto, B. Johnson, and P. Coe. 2009. Influence of cow age on grazing distribution in a mixed-conifer forest. Rangeland Ecol. Manage. 62:290-296. Walker, J. W. 1995. Viewpoint: Grazing management and research now and in the next millennium. J. Range Manage. 48:350-357. WallisDeVries, M. F. and C. Daleboudt. 1994. Foraging strategy of cattle in patchy grassland. Oecologia 100:98106. WallisDeVries, M. F., E. A. Laca, and M. W. Demment. 1999. The importance of scale of patchiness for selectivity in grazing herbivores. Oecologia 121:355363. Wang, W. and Z. Li. 2005. The distribution of grazing animals between two grassland resource points. Appl. Mat. Comp. 169:1395-1404. Warburton, K. and J. Lazarus. 1991. Tendency-distance models of social cohesion in animal groups. J. Theor. Biol. 21:473-488. Wark, T., C. Crossman, P. Valencia, D. Swain, G. BishopHurley, and R. Handcock. 2009. Pervasive Comp. IEEE. 8:30-36. Washington-Allen, R. A., T. G. Van Niel, R. D. Ramsey, and N. E. West. 2004. Remote sensing-based piosphere analysis. GISci. Remote Sens. 41:1548-1603. Weber, K. T. 2006. Challenges of integrating geospatial technologies into rangeland research and management. Rangeland Ecol. Manage. 59:38-43. Weber, K. T., M. Burcham, and C. L. Marcum. 2001. Assessing independence of animal locations with association matrices. J. Range Manage. 54:21-24. Whitehead, H. 2009. SOCPROG programs: analyzing animal social structures. Behav. Ecol. Sociobiol. 63:765-778. Wiegand, T., K. Wiegand, and S. Pütz. 2008. Grazing models. Encyclopedia Ecol. 3:1773-1782. Wikelski, M., R. W. Kays, N. J. Kasdin, K. Thorup, J. A. Smith, and G. W. Swenson, Jr. 2007. Going wild: What a global small-animal tracking system could do for experimental biologists. J. Exp. Biol. 210:181-186. Williams, R. E. 1954. Modern methods of getting uniform use of ranges. J. Range Manage 7:77-81. Williams, W. T. and P. Gillard. 1971. Pattern analysis of a grazing experiment. Aust. J. Agric. Res. 22:245-260. Willms, W., A. W. Bailey, and A. McLean. 1980. Effect of burning or clipping Agropyron spicatum in the autumn on the spring foraging behaviour of mule deer and cattle. J. Appl. Ecol. 17:69-84. Wing, M. G., A. Eklund, and L. D. Kellogg. 2005. Consumer-grade global positioning system (GPS) accuracy and reliability. J. For. 103:169-173. Wilson, R. P., W. S. Grant, and D. C. Duffy. 1986. Recording devices on free-ranging marine animals: Does measurement affect foraging performance? Ecol. 67:1091-1093. Wise, S. 2002. GIS Basics. Taylor & Francis. New York, NY. Wolff, J. O. 1993. What is the role of adults in mammalian juvenile dispersal. Oikos 68:173-176. Wong, D. W., and J. Lee. 2005. Statistical analysis of geographic information with ArcView GIS® and ArcGIS®. John Wiley and Sons, Inc. Hoboken, NJ. Workman, J. P. and J. F. Hooper. 1968. Preliminary economic evaluation of cattle distribution practices on mountain rangelands. J. Range Manage. 21:301-304. Worton, B. J. 1987. A review of models of home range for animal movement. Ecol. Model. 38:277-298. Worton, B. J. 1989. Kernel methods for estimating the utilization distribution in home-range studies. Ecol. 70:164-168. Wu, H-i., B. L. Li., T. A. Springer, and W. H. Neill. 2000. Modeling animal movement as a persistent random walk in two dimensions: Expected magnitude of net displacement. Ecol. Model. 132:115-124. Young, B. A. 1966. A simple method for recording of jaw movement patterns. J. Inst. Anim. Tech. 17:20-21. Zar, J. H. 1974. Biostatistical analysis. Prentice Hall, Inc., Englewood Cliffs, NJ. 81 Zimmerman, J. W. and R. A. Powell. 1995. Radiotelemetry error: Location error method compared with error polygons and confidence ellipses. Can. J. Zool. 73:11231133. Zuo, A. and M. S. Miller-Goodman. 2003. An index for description of landscape use by cattle. J. Range Manage. 56:146-151. 82 83 Distance traveled Temporal foraging Distance traveled Location and movement Group domestic animal activity Remote photography; animal borne video Foraging Information available to animal trackers Activities 4 5 6 7 8 Foraging location Activity and inactivity Food choices Animal location Video face detection Animal location Bite placement Animal identification Activity and inactivity Bite counter collars Telemetry system Steps during foraging Lying, standing, and walking Tracking 13 14 15 16 17 18 19 20 21 22 23 24 25 26 12 11 10 9 Walking 3 Videorecording Accelerometer; IceTag® (IceRobotics, Edinburgh, UK) UNETracker GPS/GIS/electronic compass Developed by National Agricultural Research Center, Hokkaido Pitch-roll sensor Radio frequency identification (RFID) Video/acoustic Global positioning system (GPS) Gregorini et al. (2009) Darr and Epperson (2009); Trénrl et al. (2009) Trotter (2009) Gilsdorf et al. (2008) Hiroaki and Takaaki (2007) Umstatter et al. (2006) Roberts (2006) Griffiths et al. (2006) Clark et al. (2006) Burghardt and Ćalić (2006) Ropert-Coudert and Wilson (2005) “Smart dust” Kanade-Lucas-Tomasi method Beringer et al. (2004) Müller and Schrader (2003) Porath et al. (2002); Agreil and Meuret (2004) Harris et al. (2002) Blake et al. (2002) Pedersen and Pedersen (1995) Cutler and Swann (1999); Robert et al. (2009); Moll et al. (2009) Gibb et al. (1999) Kao et al. (1995) Anderson and Urquhart (1986) Allden (1962); Stobbs (1970); Lathrop et al. (1988) Dwyer (1961) Animal mounted wireless video camera Actiwatch® Observation Night-vision binoculars Field computer Passive infrared detectors A review paper covering various types of photography; video plus other sensors Jaw movement sensor Touch panel position sensor (TPPS) Pedometers Vibracorders Following animals in a vehicle Series of switches; accelerometers Canaway et al. (1955); Robert et al. (2009) Castle et al. (1950); Moreau et al. (2009) Grazing behavior 2 Infrared equipment; tri-axial accelerometer Cory (1927) Citation(s) Table 1. Equipment for monitoring animal movement and (or) foraging No. What is recorded Sensing system 1 Distance traveled Following animals on foot 84 Water Mineral Stocking rate and season “Family groupings” Supplementation Shade Mowing Direction of wind Herding Nitrogen fertilization Time of day Distribution of experimentally altered animals may differ between collection and noncollection periods 2 3 4 5 6 7 8 9 10 11 12 Size of paddock Class of animal Plant community composition Burning Herd leadership External parasites Animal groupings and fences Herbicides Supplement timing Tree overstory Rotating animals to a new paddock Standing crop Topographic features 15 16 17 18 19 20 21 22 23 24 27 26 25 Barometric pressure and wind speed 14 13 Salt 1 Items such as brush, cliffs, and ravines affect distribution Previous year‟s dead standing crop affects current year use. Movement of cows to a new paddock increased mean daily travel 30%. Horn flies increased cattle travel. Sterile bulls spent more time closer to fences and corners than steers when heifers were present in an adjoining paddock. Cows preferred foraging on plots treated with 20% tebuthiuron pellets compared to untreated plots regardless of grass species in the plots. Steer supplemented in the AM tended (P < 0.2) to graze longer than steers supplemented in the PM while steers receiving no supplement traveled less (P < 0.02) than either AM or PM supplemented steers. Cattle prefer foraging under open canopied forest compared to a dense canopy. Certain cows may initiate movement and direction affecting the entire herd. Yearling cattle can travel further than pairs. Forage consumption of yearling cattle increased as grass production increased and forest density decreased. Cattle preferred burned to clipped or control treatments of Agropyron spicatum. Time spent foraging and ruminating increased in response to either a rising or falling pressure but these changes were short lived while distance travel was inversely correlated to mean wind velocity. Subdividing large paddocks into smaller units improves distribution. Esophageally fistulated cattle may forage closer to fences and corrals during sampling compared with periods of non-sampling. Cattle foraged during the night and loafed around drinking water during the day. Animals are drawn to green forage. Optimizing spatial and temporal decisions based on experience. Artificial shade can alter cattle dispersion. Cattle preferred mowed brush treatments, especially in the first year following the treatment. Cattle and horses graze with the wind; sheep and goats graze into the wind. Supplemented cows walked less than non-supplemented cows. Strategic placement of salt. Cattle under heavy stocking compared with light stocking exerted more “vigorous” foraging with total length of daylight foraging increasing from August to October. Certain animals prefer to remain in the presence of known peers. Forage utilization decreases as distance from drinking water increases. Will aid in the distribution of free-ranging cattle. Table 2. Various attributes that have been shown to affect animal distribution with selected references No. Attribute Observation Selected references Senft et al. (1987) Norton and Johnson (1986) Anderson and Urquhart (1986) Mitchell and Rodgers (1985) Adams (1985) Scifres et al. (1983) Hinch et al. (1982) Harvey and Lauchbaugh (1982) Reinhardt and Reinhardt (1981) Willms et al. (1980) Clary et al. (1978) Arnold and Dudzinski (1978) Sato et al. (1976); Hunt et al. (2007) Malechek and Smith (1976) Dean and Rice (1974) Workman and Hooper (1968); East (2003) Hooper et al. (1969); Smith and Lang (1958); Samuel et al. (1980) Schmidt (1969) Hafez (1968); Squires (1978) Powell and Box (1966) McIlvain and Shoop (1965) Box et al. (1965) Skovlin (1957) Peterson and Woolfolk (1955) Williams (1954) Chapline and Talbot (1926); Martin and Ward (1973) Valentine (1947); Franklin et al. (2009) 85 Memory Herd size Drought Animal memory Attention of the livestock manager Slope Supplement and walking Size of animal groupings Fencing Body size Patch size and condition Interspecific socialization Early life experiences Historical foraging patterns Selection Trails Visual cues Animal experience Supplement placement Diurnal patterns Shape of paddock Size of preferred patch Domestic stock may affect wildlife distribution Fire Integration of several factors simultaneously Kind of animal 29 30 31 32 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 54 53 52 51 34 33 Rotation schedules 28 Cattle travel further from water than goats, which travel further than sheep while feral pigs are more reliant on access to water than livestock and rabbits are not dependent on surface water. Fires may promote more uniform grazing following a burn. Season of use, supplement placement, drinking water development, and herding managed simultaneously can positively affect distribution. Mule deer and elk avoided areas used by cattle. Yearling steers could associate visual cues with feed quality to forage more efficiently Animals unfamiliar with an area widen their searching leading to a more uniform distribution during foraging compared with experienced animals. Improved utilization from 0 to 600 m from supplement location during late summer, autumn, and winter. As daylight decreases, night time foraging tends to increase. Sheep tend to utilize herbage more efficiently in square rather than rectangular pastures; similar data does not exist for cattle. Heifers and ewes preferred larger patches over smaller patches of ryegrass. Current foraging programs seldom consider the previous years utilization Steers apparently use different processes when selection between compared to selecting within feeding stations. Cattle prefer to use trails. Small and medium sized cows were more selective in their diets compared with males. Steers selected short and tall patches of Agrostis/Festuca and Lolium grasslands over stemmy patches but preferred short patches. Bonding of small ruminants (sheep and goats) to cattle will alter the spatial distribution of the small ruminants in free-ranging conditions. Cattle tend to return to areas in which they were raised as calves. Creating smaller paddocks forces more use of the entire area. Cows remember locations and previously selected diets. With the implementation of grazing systems more attention to all aspects of management may increase, including those affecting distribution. Cattle avoid slopes greater than 7%. Steers supplemented with cotton seed meal (CSM) did not spend more time walking than unsupplemented steers grazing wheatgrass pasture. Bison calves followed dams more when in small groups than in larger groups Cattle forage more evenly when rotation schedules were timed to plant growth not fixed calendar dates. Past foraging locations may be remembered by cattle up to 8 h. Behavior in a herd of 3 cattle is likely to be atypical when herd size varied between 3 and 24 animals. Cattle tend to disperse. Table 2 (Continued). Various attributes that have been shown to affect animal distribution with selected references No. Attribute Observation Pringle and Landsberg (2004) Bailey (2004) Archibald and Bond (2004) Stewart et al. (2002) Dumont et al. (2002) Sevi et al. (2001); Laca (2009) Linnane et al. (2001) Bailey et al. (2001) Ksiksi and Laca (2000) Howery et al., 2000 Ganskopp et al. (2000) WallisDeVries et al. (1999) Vavra and Ganskopp (1998) Howery et.al. (1998) Anderson (1998) WallisDeVries and Daleboudt (1994) Lazo and Soriguer (1993) Hart et al. (1993) Green (1992) Barton et al. (1992) Pinchak et al. (1991) Hart et al. (1991) Bailey and Rittenhouse (1989) Pick and Chewings (1988) Hacker et al. (1988) Bailey (1988) Anderson (1988) Selected references 86 Breed of animal Intensive vs. extensive management Conspecific attraction Man made structures Visual cues Coat color Socioeconomic factors 57 60 61 62 63 Cow age (experience) Use of new management technologies 68 Conspecifics 67 66 65 64 59 Seasonal travel and breed differences Season Cow phenology 56 58 Standing crop composition 55 Cattle tend to group up when attempting to cross a virtual fence line. Cattle used roads as paths of least resistance. Heifers grazed more high-quality patches compared to low quality patches when visual cues were placed in those patches. Holstein Friesian dairy cows with predominantly black coats used shade less than cows with predominately white or black and white coats. Government regulations may prevent producers from implementing ecologically optimal management strategies. Cows decreased daily walking even as herbage quality declined between spring and summer; Baladi cows walked longer daily distances than Beefmaster x Simford crosses. Heifer foraging time increased between June and September. Naïve calves showed a higher foraging activity on their first day on pasture when in the presence of experienced conspecifics compared to naïve calves left on their own. Cows greater than 3 years-of-age preferred areas of less canopy closure; young cows selected lower elevation and steeper slopes compared with the oldest cows. Cattle prefer to forage on standing crop that has previously been grazed (green) compared to foraging in areas with both green and cured material, suggesting utilization patterns are self-sustaining. Cows that were heavier and taller traveled further (horizontally) from water than did lighter cows. Spanish (Corriente) cattle travel further than Angus and Herefords. Cattle managed under intensive management foraged more evenly over the landscape compared to cattle managed under extensive management. An unexplored potential mechanism possibly affecting home range behaviors. Table 2 (Continued). Various attributes that have been shown to affect animal distribution with selected references No. Attribute Observation Wark et al. (2009) Walburger et al. (2009) Hessle (2009) Hejcmanová et al. (2009) Aharoni et al. (2009) Wiegand et al. (2008) Tucker et al. (2008) Renken et al. (2008) Cooper et al. (2008) Börger et. al. (2008) Barnes et al. (2008) Sheehy (2007) VanWagoner et al. (2006) Ganskopp and Bohnert (2006) Selected references 87 Tracking Analyst BIOTAS Excel HOME RANGE* LOAS LOTE Lotus 1-2-3 MacComp MARK MATLAB NCSS QuattroPro RANGES 8 5 6 7 8 9 10 11 12 13 14 15 16 17 Operating system D W W M W D W W W W D W A M M D D F M x x x x x x x x x x x x Data formatting and error checking Autocorrelation x x x x x x x x Travel path Animal interactions x Kernel F/A F/A F F/A F/A A X x x x x x x x x x x x x x x x x x X X X x x x x X x Harmonic mean Bivariate normal Fourier Minimum Convex Polygon x x x x http://www.anatrack.com/ http://www.esri.com/software/arcview/index.html http://www.absc.usgs.gov/glba/gistools/ http://www.esri.com/software/arcgis/extensions/spatia lanalyst/index.html http://www.esri.com/software/arcgis/extensions/tracki nganalyst/index.html http://www.ecostats.com/software/biotas/biotas.htm http://www.microsoft.com/office/excel/default.htm http://www.cnrhome.uidaho.edu/default.aspx?pid=730 83 http://www.ecostats.com/software/loas/loas.htm http://www.ecostats.com/software/lote/lote.htm http://www.lotus.com/home.nsf/tabs/lotus123 http://detritus.inhs.uiuc.edu/wes/habitat.html http://www.cnr.colostate.edu/~gwhite/mark/mark.htm http://www.mathworks.com http://www.ncss.com http://www.corel.com http://detritus.inhs.uiuc.edu/wes/home_range.html Source as of November 18, 2009 Adapted from Larson (2001) KEY: HOME RANGE for use in ARC/INFO, which is a product of ESRI, Redlands, CA USA. *Operating System: A = indicates that some sample code is available to be compiler for the designed operating system, D = DOS, F = Fortran, M = Macintosh, W = Microsoft Windows Travel Path: includes animated graphics and analyses of angles, distances, and rates of animal movement between successive telemetry locations. Animal Interactions: includes analyses of site fidelity and static and dynamic interactions among marked individuals. Kernel: indicators the program calculates for (F)ixed, (A)dative, or both (F/A) kernels, when known. 1 Spatial Analyst Antelope Calhome Dixon Homer HomeRange* Kernelhr McPaal Home Ranger Ulysses Wildtrak ArcView and extensions Animal Movements 4 3 2 1 No. Cluster Home range estimators Grid cell Table 3. Software useful for monitoring animal movements with information available over the web 1 Digitized polygon 88 19 18 Animal-habitat relationships Distribution during foraging and utilization Animal movement in twodimensions Utilization of space 16 17 Home range 15 11 12 13 14 10 9 8 7 Vegetation types and distance from water Utilization distribution (UD) Movement patterns Spatial and temporal utilization on foothill range Response of animals to their environment Habitat use Cattle home range size Spatial distribution Distribution Animal movement 5 6 Orientation and movement Grazing + travel, resting, and bedding patterns Real-time duration and latency measures Animal activity area 4 3 2 1 None specified None specified 30 None specified None specified None specified 1200 2,430 None specified None specified 4,597 248 None specified 170,000 None specified None specified None specified 11-125 None specified Minimum convex polygons (MCP), fixed-kernels and clusters Model of persistent 2D random walks Tobit analysis; Proc Genmod Genetic algorithm for rule set production (GARP) Log-ratio analysis of compositions CALHOME Multivariate general linear modeling Convection-diffusion approach Kernel Home Range program (KERNELHR) Nonparametric model in which the Thiessen model delineated areas of intensive use more effectively than Kernel estimators Chi-square analyses Transition matrix and cluster analysis Kernel methods Convection-diffusion equation involving the inverse Gaussian density function Biased random walk movement model Multifactorial sets of angular data Kruskal-Wallis & Mann-Whitney statistical analysis Stepwise multiple regression Harmonic mean of an aerial distribution Table 4. Mathematical approaches to evaluate animal distribution and its associated factors Area No. Parameter(s) Investigated Analytical Tool (ha, range) Wu et al. (2000) Girard et al. (2002) Mohr (1947); Worton (1989); and Kenward (1987) Brock and Owensby (2000) Casaer et al. (1999) Seaman et al. (1998) Aebischer et al. (1993) Kie and Boroski (1996) Landsberg and Stol (1996) Pickup and Bastin (1997) Stockwell and Noble (1992) Pinchak et al. (1991) Bailey et al. (1990) Worton (1989) Pickup and Chewings (1988) Marsh and Jones (1988) Underwood and Chapman (1985) Hendrie and Bennett (1984) Senft et al. (1983) Dixon and Chapman (1980) Manuscript Citation Wu et al. (2000) Tobin (1958); SAS (1993) Boots (1980) Worton (1989) Stockwell and Noble (1992) Aitchison (1986) Kie et al. (1996) SYSTAT (1992) Pickup (1994) Zar (1974) Lehner (1979) Worton (1989) Kareiva and Shigesada (1983; correlated random walk) Moore and Clarke (1983); Folks and Chhikara (1978) Mardia (1972) Hendrie and Bennett (1984) None provided Neft (1966) Analytical Tool Citation 89 39 38 37 35 36 34 33 32 31 30 29 28 27 25 26 24 23 Group interactions Distribution Movement based on object‟s trajectory Temporal and spatial location of cattle Activity vs. inactivity Distance traveled Landscape use Distribution between two resource points Animal spatial and temporal groupings Grazing, traveling, and resting Tracking distribution changes Time spent foraging and distance traveled Animal dispersal Resting, grazing, and walking Distances between individuals Foraging paths Foraging Detection of animals 22 21 Feces pattern Distribution Evenness Index (DEI) Parameter(s) Investigated 20 No. 466 None specified 329 & 258 466 250-1530 4,168 900 - 57,000 75.2 None specified 28-859 810 2,350 0.15 - 0.35 2,000 120,000190,000 0.05 None specified None specified 3.3 150.5 Area (ha, range) Jeung et al. (2008) Ljung (1999) System identification with least-squares fitting Littell et al. (1996) Arabie et al. (1996) Estevez and Tatem (2007) Kiker et al. (2006) Rogers and Carr (2002); Worton (1989) Weber et al. (2001) Li and Wang (1998a,b) Breiman et al. (1984) and SAS (2002) Schauer et al. (2005) www.baylor.edu/grass Long et al. (2005) Milne (1992) Kenward (1987) None specified SAS (2000) Pinheiro and Bates (2000) Shannon and Weaver (1949) Ramsey and Harrison (2004) Analytical Tool Citation Hybrid Prediction Model PROC MIXED Home Range extension for ArcView; kernel analysis Integration of a habitat suitability index (HSI) into ACRU2000 K-means classification algorithm Spatial Analyst in ArcMap9.1 Association software, ASSOC1 Repeated statement with the Mixed procedure of SAS Multiple regression, discrimination analyses and classification and regression trees together with the JMP statistical analysis program (Release 5.01a) Stochastic process and differential equation GRASS software Meta-analysis Fractal analysis Grid cell method Simple and multiple regression Discriminate analysis Distance sampling Shannon-Wiener Index Linear mixed effects model Analytical Tool Table 4 (Continued). Mathematical approaches to evaluate animal distribution and its associated factors Schwager et al. (2008) Jeung et al. (2008) Bailey et al. (2008) Schwager et al. (2007) Tomkins and O‟Reagain (2007) Pandey (2007) Hunt et al. (2007) Harris et al. (2007) Wang and Li (2005) Ungar et al. (2005) Schauer et al. (2005) Schacht et al. (2005) Long et al. (2005) Garcia et al. (2005) Kawamura et al. (2005) Shiyomi (2004) Schlecht et al. (2004) Ramsey and Harrison (2004) Tate et al. (2003) Zuo and Miller-Goodman (2003) Manuscript Citation 90 Table 4 (Continued). Mathematical approaches to evaluate animal distribution and its associated factors Area No. Parameter(s) Investigated Analytical Tool Analytical Tool Citation (ha, range) Calenge (2006); Object class “ltraj” developed for 40 Trajectory analysis None specified R Development Core R software Team (2006) 41 Foraging / walking 829-864 Forward stepwise regression Ungar et al. (2005) Herd movement using Hidden Markov models and long-term Bharucha-Reid (1960); 42 individuals as the central 7 trajectory prediction Murphy (1998) unit of measure Observer Kalman filter identification Jung (1994; OKID); Pitch angle of neck (animal 43 None specified (OKID) and multiple-model adaptive Ferreira and Waldmann activity and inactivity) estimation (MMAE) (2007) Log-transformed/residual REML; Genstat 8.2, VSN 44 Distance traveled 93-117 maximum likelihood International, Hertz, UK 45 Animal associations None specified SOCPROG Whitehead (2009) Whitehead (2009) Tomkins et al. (2009) Nadimi and Søgaard (2009) Guo et al. (2009) Ganskopp and Bohnert (2009) Calenge et al. (2009) Manuscript Citation GEOSPATIAL METHODS AND DATA ANALYSIS FOR ASSESSING DISTRIBUTION OF GRAZING LIVESTOCK D. M. Anderson Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 METABOLIC SIGNALS OF THE BEEF COW IN NEGATIVE ENERGY BALANCE1 R. C. Waterman2ŧ and W. R. Butler* ŧ USDA - ARS, Fort Keogh Livestock and Range Research Laboratory, Miles City, MT 59301; * Department of Animal Science, Cornell University, Ithaca, NY 14853 grasses (Heitschmidt et al., 1995, 1999; Grings et al., 2005). As summer progresses, temperatures increase while precipitation decreases, followed by decreases in forage quality that continue through fall and winter, which negatively impacts livestock production (Adams and Short, 1988). Waterman et al. (2006) demonstrated that varying concentrations of glucogenic precursors fed to postpartum range cows consuming supplements containing glucogenic AA with or without Ca-propionate influenced nutrient partitioning. Supplements with greater concentrations of glucogenic precursors minimized maternal tissue catabolism and allowed partitioning of dietary nutrients toward maternal tissue. As a result, rates of glucose sequestration into maternal tissues were modified. If nutritional stresses that render maternal tissues less responsive to actions of insulin can be identified, then metabolically targeted supplementation protocols may be designed to minimize this effect. Seasonal alterations in glucose metabolism (i.e., oxidative metabolism), as influenced by tissue responsiveness to insulin in beef cows grazing rangelands in the Northern Great Plains, occur in relation to forage quality. Altered glucose metabolism may ultimately affect physiological performance and reproductive function in range beef cows (Waterman et al., 2007). INTRODUCTION Extensive beef cattle enterprises endure difficulties trying to maintain beef cow production that is both sustainable and profitable. Furthermore, beef cow production on arid and semiarid rangelands is directly influenced by environmental conditions (i.e., forage quantity and quality available for consumption). To achieve optimal levels of production, harvested and purchased feedstuffs are often fed to range cows to supplement periods of nutritional stress. The cost of delivering these feedstuffs to range cows is one of the largest input costs a producer will encounter. Moreover, the range beef cow still must perform by being reproductively and biologically efficient within these changing environments. Under these demanding circumstances, beef cows experience similar metabolic dysfunctions as observed in high producing dairy cows following parturition; however, seasonal changes in forage quality play a larger role in beef cow production and can induce metabolic dysfunctions throughout the production cycle year at times besides around parturition. Excellent reviews that discuss the nutritional controls on reproduction are available for the beef cow (Keisler and Lucy, 1996; Hess et al., 2005) and the dairy cow (Bauman and Currie, 1980; Butler, 2000; van Knegsel et al., 2005; Chagas et al., 2007; van Knegsel et al., 2007). The purpose of this proceedings paper will be to expand the topic of negative energy balance that occurs around parturition and (or) early lactation to include the entire production cycle of the beef cow. DERIVED METABOLIC SIGNALS AND TARGETED TISSUES OF BEEF COWS IN NEGATIVE ENERGY BALANCE In order to better understand the intertwined web of metabolic signals, it is essential to review the tissues involved and the metabolic signals they elicit. Therefore, the initial portion of this proceeding is designed to describe specific tissues involved in producing metabolic signals and the targeted tissues of those signals when beef cows encounter negative energy balance. NUTRITIONAL ENVIRONMENT INFLUENCING BEEF COWS IN NEGATIVE ENERGY BALANCE In extensive western livestock production systems, timing and amount of precipitation are crucial in spring through early summer for establishment of cool-season Rumen 1 USDA, Agricultural Research Service, Northern Plains Area, is an equal opportunity/affirmative action employer, and all agency services are available without discrimination. Mention of any trade name or proprietary product does not constitute a guarantee or warranty of the product by USDA, Montana Agric. Exp. Stn., or the authors and does not imply its approval to the exclusion of other products that also may be suitable. 2 Corresponding author: [email protected] Ruminants have a unique synergistic relationship with microflora that inhabits the reticlo-rumen. This relationship enables cows to consume forages high in fiber (cellulose and hemicellulose) that are fermented by the ruminal microflora to produce VFA. The ruminal microflora then uses dietary N and fermentable energy sources to manufacture microbial cells (i.e., microbial protein), which in turn, can be utilized by the host ruminant. 93 receptors in the anterior pituitary and hypothalamus via peripheral circulation (Tschop et al., 2000). In ruminants, ghrelin concentrations increase before scheduled meals (Sugino et al., 2002) or in response to fasting (Wertz-Lutz et al., 2006), and are less variable when diets are consumed ad libitum (Sugino et al., 2002). Upon feeding, ghrelin concentrations are suppressed (Wertz-Lutz et al., 2006). Other gut peptides that may play pivotal roles during negative energy balance include gastrin, cholecystokinin (CCK), glucose-dependent insulinotropic polypeptide (GIP), and glucagon-like peptide 1 (GLP-1). The exact role that these gut peptides play in ruminant metabolism is unclear. However, Relling and Reynolds (2007) speculated that these peptides may play a role in “fine-tuning” nutrient metabolism through their anabolic effects on adipose or mammary gland metabolism. Concentrations of GIP and GLP-1 were greater in lactating sheep than in non-lactating sheep (Faulkner and Martin, 1997). Insulin concentrations were lesser in lactating ewes than in non-lactating ewes, indicating a lack of association between GIP, GLP-1, and insulin. Gastrin appears to stimulate pancreatic release of insulin and glucagon as dietary nutrients reach the small intestine (Bloom et al., 1975; Brockman, 1978). The primary VFA produced by the ruminal microflora include acetate, butyrate, and propionate. Beef cows confined to open rangelands are reliant upon environmental conditions and often consume senescent forages leading to periods of inadequate nutrition (Hawkins et al., 2000). During these conditions, beef cows experience high ruminal acetate production, but propionate production is often considered to be inadequate. Propionate is important as the principle glucogenic precursor in ruminants (Bergman, 1973; Forbes, 1982; Bell and Bauman, 1997) and when lacking, other glucogenic substrates (i.e., glucogenic AA, glycerol, and lactate) derived from maternal catabolism (i.e., loss of body reserves) are required for gluconeogenesis. Acetate is primarily absorbed through the rumen epithelium virtually unchanged with very little conversion to ketones. However, ruminal butyrate is largely converted to ketones (β-hydroxybutyric acid, acetoacetate, and acetone) by the rumen epithelium. Ruminal propionate also passes primarily unchanged through the ruminal epithelium with only small amounts being converted to lactic acid (Fahey and Berger, 1987). When cellular energy from glucose is compromised, the secondary energy source becomes ketones. Volatile fatty acids absorbed through the ruminal epithelium enter the peripheral blood supply and are predominately taken up by the liver. Ketones formed from ruminal VFA across the rumen epithelium enter the peripheral blood supply and can be used in oxidative metabolism by heart, kidney, skeletal muscle, and the lactating mammary gland. However, ketone usage by the ruminant brain may be minimal (Pell and Bergman, 1983). Another important metabolic signal originating in the rumen includes various forms of N. Ruminant animals have the unique ability to conserve N or optimize N balance in the rumen. Excess ruminal N is transported across the ruminal epithelium as ammonia by simple diffusion of the non-ionized lipid soluble form [NH3; (McDonald, 1948)] or NH4+ via potassium channels (Abdoun et al., 2006). Ammonia absorbed into the mesenteric and portal blood supply is removed by the liver and converted to urea and glutamine (Parker et al., 1995). During inadequate nutritional conditions, N intake deficiencies reduce ruminal microbial cell synthesis, DMI, and energy availability, thus resulting in diminished fermentation (Hawkins et al., 1999). Fundamental Principles: Metabolic signals derived from the gastrointestinal tract include ghrelin, gastrin, CCK, GIP, and GLP-1. The metabolic roles of these signals are influenced by nutrients presented to the stomach and small intestine. Liver As stated by Forbes (1982), “The liver is at the crossroads of metabolism”. The liver is responsible for detoxification (i.e., ammonia to urea), production of cellular fuel (i.e., glucose) via gluconeogenesis, synthesis of ketones, production of insulin-like growth factors (IGFs; which are polypeptides with conserved sequences and actions similar to insulin), and storage of glycogen. An in depth review of N metabolism by the liver is beyond the scope of this proceeding; readers are directed to the review of Reynolds (1992) for more detail. As Reynolds (1992) pointed out, a substantial amount of dietary N is absorbed via the rumen and transported to the liver as ammonia. The role of the liver is to detoxify ammonia primarily by converting it to urea, which is then released into peripheral circulation to be recycled back into the rumen via saliva or direct absorption across the ruminal wall via a concentration gradient. Excess urea not recycled back to the rumen is excreted to the environment through urine. Another role of N in the liver is utilization of peripheral N in the form of glucogenic AA for glucose synthesis (i.e., gluconeogenesis). Although this is an inefficient use of N, it is essential, especially for ruminants exhibiting negative energy balance and when ruminal fermentation is unable to provide an adequate supply of propionate. Propionate is the principle glucogenic precursor in ruminants (Bergman, 1973; Forbes, 1982; Bell and Bauman, 1997) and when inadequate, other glucogenic substrates (i.e., glucogenic AA, glycerol, and lactate) are Fundamental Principles: Metabolic signals derived from the rumen include ammonia (processed in the liver to urea and important in N recycling back to the rumen) and the VFA, acetate and butyrate (primarily converted to ketones, β-hydroxybutyric acid, acetoacetate and acetone), and propionate (primarily processed in the liver and important for proper TCA cycle function and gluconeogenesis). Gastrointestinal Tract Ghrelin is a 28-amino acid peptide secreted primarily by abomasal and ruminal cells (Hayashida et al., 2001; Bradford and Allen, 2008) that stimulates appetite and secretion of growth hormone (GH). Ghrelin acts on 94 utilized, resulting in the catabolism of maternal tissues (loss of body weight) in an attempt to satisfy the energy requirement of a range beef cow. Also, metabolizable protein (e.g., ruminally undegradable protein and microbial protein) absorbed from the small intestine may be inadequate to properly support gluconeogenesis. Thus, concurrent with inadequate gluconeogenesis, the tricarboxylic acid cycle becomes unable to regenerate oxaloacetate (OAA), which is a key intermediate required for gluconeogenesis (Salway, 1994) and acetate metabolism in the liver. This results in the utilization of acetate in futile cycles, which increases heat production and decreases energetic efficiency while redirecting acetate carbon towards the synthesis of ketones. As stated previously, this alteration in metabolism is exaggerated by a decrease in DMI, which reduces the supply of incoming ruminal acetate, and a reduction in the efficiency of energy metabolism as reflected by increases in serum ketones, which are primarily in the form of β-hydroxybutyrate (Kaneko, 1989). During negative energy balance in cattle, insulin-like growth factor 1 (IGF-1) becomes uncoupled from GH stimulation, and as a result, the liver produces less IGF-1 (Lucy et al., 2009). This uncoupling between GH and IGF-1 contributes to delayed return to ovulatory cycles and prolonged anestrous. Another role of IGF-1 is the insulinlike anabolic actions it initiates in other tissues. Muscle In times of negative energy balance, muscle responds to endocrine systems by catabolizing protein rich tissues (Chagas et al., 2007). As stated above, glucogenic AA are important precursors for gluconeogenesis; however, a balance must be maintained to avoid excessive gluconeogenesis at the expense of peripheral protein by sparing glucose utilization (Brockman, 1978). During periods of low energy availability, muscle produces myostatin, a member of the transforming growth factor-β superfamily (Lee, 2004). Myostatin promotes mobilization of muscle proteins as free AA, making them available for use by other tissues (Zimmers et al., 2002). Increased myostatin in the blood stream seems to down regulate the insulin-dependent glucose transporter, Glut 4, thus increasing insulin resistance by muscle and adipose tissue (Antony et al., 2007). Fundamental Principles: Metabolic signals derived from muscle include AA and myostatin. These signals are derived when muscle catabolism occurs. Adipose Tissue Adipose tissue provides a crucial role during undernutrition of ruminants (Chilliard et al., 2000). When ruminants experience undernutrition, glucagon exerts its effects on adipose tissue by stimulating hormone-sensitive lipase, thereby activating lipolysis and converting triglycerides into glycerol and NEFA, which are released into the peripheral circulation to be used for energy by other tissues. Non-esterified fatty acids are metabolized by the liver to contribute acetyl CoA to the TCA cycle, but also result in the formation of ketones when OAA cannot be efficiently regenerated. Chilliard et al., (2000) indicates that apart from adipose tissue serving as an energy storage site, it acts as an endocrine gland secreting autocrine, paracrine, and endocrine factors. Leptin secretion by adipose tissue is markedly reduced during undernutrition in ruminants, resulting in neuroendocrine inhibition of the satiety control center to encourage DMI. Conversely, this neuroendocrine stimuli may also have implications that extend anestrous via inhibition of the reproductive axis (Friedman and Halaas, 1998; Heiman et al., 1999). Fundamental Principles: Metabolic signals derived from the liver include urea, glucose, ketones, and IGF1. These signals are influenced by the peripheral concentrations of circulating precursors. Pancreas The pancreas has an important role in animal metabolism; however, for purposes of this proceeding we will focus on the roles of insulin and glucagon. The pancreas also has other important exocrine roles and the authors direct the reader to the review by Croom et al., (1992) for more details. Insulin and glucagon play important roles in regulating intermediary metabolism. Insulin (hypoglycemic hormone) promotes glucose uptake from blood by peripheral tissues (Bowen, 1964; West and Passey, 1967), including muscle (Jarrett et al., 1974) and adipose tissue (Khachadurian et al., 1966), and promotes storage of other metabolites in peripheral tissues (Brockman, 1978). Ruminants are different than nonruminants such that gluconeogenesis and lipogenesis occur simultaneously via substrates derived from ruminal fermentation (VFA; propionate and acetate, respectively). Glucagon (hyperglycemic hormone) promotes both gluconeogenesis and lipolysis to maintain glucose homeostasis (Brockman, 1978). Fundamental Principles: Metabolic signals derived from the adipose tissue include NEFA and leptin. These signals are important in energy status. Uterus and Ovary Ovarian and normal estrous cycle activities are suppressed during undernutrition or when cows are experiencing negative energy balance. Furthermore, negative feedback to the hypothalamus via estradiol extends the anestrous period in cows experiencing negative energy balance (Short et al., 1990; Keisler and Lucy, 1996) and compounds the inhibitory effects of suckling by the calf. Fundamental Principles: Metabolic signals derived from the pancreas include insulin and glucagon; primary role is the regulation of glucose in the peripheral circulation. Muscle 95 produces catecholamines (i.e., epinephrine) which are also responsive to ACTH. Epinephrine elicits a response at the liver to increase gluconeogenesis, muscle to mobilize protein, and adipose to mobilize triglycerides and fatty acids. Obviously, an extended period of anestrous and delayed ovulation reduces the reproductive capacity of range beef cows during the breeding period. Fundamental Principles: Metabolic signals perceived by other tissues can be exacerbated by ovarian estradiol. Estradiol signaling is essential to achieve post-partum reproductive competency. Fundamental Principles: Metabolic signals derived from the adrenal gland include cortisol and epinephrine. These signals are important in energetic efficiency. Brain (Hypothalamus and Pituitary) INTERTWINED WEB OF METABOLIC SIGNALS FOR BEEF COWS IN NEGATIVE ENERGY BALANCE The satiety control center located in the ventromedial hypothalamic nucleus is responsible for modulating signals of hunger (orexigenic) and satiety (anorexigenic) originating from numerous peripheral tissues throughout the ruminant (Roche et al., 2008). The authors refer readers to the review by Roche et al., (2008) for greater detail on the neuroendocrine and physiological regulation of intake in ruminants. Neuropeptide Y is a 36-amino acid peptide synthesized in both the central nervous system and the peripheral nerves (Hulshof et al., 1994). Animals that are nutritionally compromised have elevated levels of NPY indicating that one role of NPY is to stimulate feed intake (McShane et al., 1993; Prasad et al., 1993; Thomas et al., 1999). Additionally, the preoptic area (POA) of the hypothalamus is responsive to neuronal signals (NPY and opiates), reproductive hormones (E2 and P4), as well as nutritional metabolites (glucose, acetate, ketones, NEFA, and excitatory AA). The medial basal hypothalamus (MBH) is responsive to signals from leptin, insulin, and IGF-1. Neuroendocrine responses from the POA and MBH regulate anterior pituitary secretion of follicle stimulating hormone and luteinizing hormone (FSH and LH, respectively) that mediate and control ovarian follicular development necessary for ovulation. Lastly, during parturition when the fetus is presented to the cervix, signals are sent to the hypothalamus that then signals the posterior pituitary to release oxytocin. Oxytocin influences the mammary gland to initiate milk let down. This action prompts mammary tissue to sequester glucose, which creates a nutrient sink that exacerbates the requirement for glucose, especially when gluconeogenesis is compromised due to the lack of glucogenic precursors. Parturition and (or) Early Lactation The metabolic conditions occurring in range beef cows that ultimately lead to negative energy balance are in principle the same as those observed for high producing dairy cows at parturition and through early lactation. The exception is that high producing dairy cows have been genetically selected for high milk production, which directly sets up a cascade of events leading to negative energy balance (and possibly ketosis) just before parturition that may continue through early lactation. Range beef cows are less likely to exhibit negative energy balance or ketosis directly from the additional demands for lactation, but rather are driven into negative energy balance by the inability to consume enough dietary energy to support requirements for both maintenance and lactation, not to mention the requirements for growth in young developing beef cows. The dairy cow generally consumes a total mixed ration that is formulated to meet her requirements; however, the extreme demand for milk production exceeds the biological capacity for oxidative metabolism, which in turn, creates potential for the occurrence of negative energy balance. The most crucial time for dairy cows is the two weeks preceding parturition and the first few weeks of lactation. For range beef cows, this change can also originate before parturition, but negative energy balance is reached much more gradually. The endocrine changes surrounding parturition and onset of lactation have metabolic effects in multiple tissues. An increased demand for glucose requires an increase in glucogenic precursors; however, DMI is sometimes suppressed in late gestation leading to a reduction in fermentable substrate to support gluconeogenesis (i.e., decreased propionate). Liver glycogen stores are most likely depleted due to the dam’s increased energy requirements leading up to and during parturition. Therefore, glucogenic precursors must come from catabolism of maternal tissues. Initially, stress hormones (i.e., cortisol and epinephrine) target adipose tissue and stimulate lipolysis, which releases triglycerides and NEFA into the peripheral circulation. These hormones also target muscle tissue to initiate protein catabolism releasing AA into the blood stream. Another role for epinephrine is to target the liver and stimulate gluconeogenesis. The catabolism of muscle mobilizes glucogenic AA to the liver, but this supply may not be sufficient to regenerate Fundamental Principles: Metabolic signals derived from the brain (hypothalamus and pituitary) include GnRH, GHRH, CRF, ACTH, LH, FSH, GH, oxytocin, and NPY. These signals are essential for energetic and reproductive competency. Adrenal Gland The adrenal gland produces glucocorticoids (i.e., cortisol) in the adrenal cortex in response to pituitary release of adrenocorticotropic hormone (ACTH). Cortisol elicits a response at the liver to increase gluconeogenesis, muscle to mobilize protein, adipose to mobilize triglycerides and fatty acids, and consequently directly inhibits glucose transport into cells. The adrenal medulla 96 parturition. As a result, range beef cows may experience periods of short and extended negative energy balance. Undernutrition may occur during early gestation and potentially cause early embryonic mortality. Therefore, the initial cause that exacerbates maternal catabolism of tissues is the cow’s inability to obtain sufficient nutrients from their diet to meet requirements for maintenance, growth, and gestation. Once a pregnancy is established, the cow will usually provide sufficient nutrients to maintain that pregnancy at the expense of maintenance and growth. The beef cow may be able to reestablish her maintenance requirements by decreasing activity and DMI. Range beef cows can rebound quickly in the short term; however, extended periods of undernutrition can greatly impact the beef cow’s oxidative metabolism leading to undesirable performance. If the period of negative energy balance occurs in the last third of gestation, there is a greater propensity for dystocia and increased negative energy balance at parturition on into lactation. The impact of negative energy balance on in utero development of the fetus is not well understood and deserves further study. A body of new information is coming forthwith suggesting that nutrient deficiencies in utero may compromise production traits of the developing fetus. A literature review by Du et al. (2009) demonstrated that nutrient deficiencies at key times during gestation can alter not only the number of muscle fibers but their growth as well. The metabolic signals associated with negative energy balance are very similar at any time throughout the production cycle, with the main differences being the removal of signals produced or directed to reproductive organs once pregnancy is established. As range beef cows begin to experience negative energy balance, stress hormones from the adrenal gland are secreted along with pancreatic release of glucagon in an effort to maintain homeorhesis. These events can cascade, and if sustained for a lengthy period of time, the cow experiences severe negative energy balance. This is again caused by the lack of OAA regeneration in the TCA cycle due to insufficient supply of glucogenic precursors. The condition is manifested by the inability of insulin to elicit an effect on tissues and allow the translocation of the Glut 4 transporter to the cell membrane for glucose and other essential nutrient uptake into tissue cells. Dietary nutrients must be provided in sufficient quantities to repartition nutrients toward maternal tissue while meeting the needs of the animal for maintenance, gestation, and growth in younger cows. Harrelson et al., (2009) compared feeding a traditional cottonseed meal supplement [36% CP (35% ruminally undegradable protein)] at 454 g/d and prorated to be delivered 3 d/wk compared to a biologically potent small package supplement [33% CP (60% ruminally undegradable protein)] offered daily at 113 g/d. Cows receiving the small package supplement experienced rapid glucose clearance from blood, indicating faster uptake of glucose by tissues. This would also stipulate that ketones and NEFA concentrations in the peripheral blood supply were reduced with the small package supplement, a result of faster glucose clearance rates and improved oxidative metabolism (i.e., regeneration of OAA). OAA in the TCA cycle fast enough to utilize the excess acetyl-CoA supplied to the liver by NEFA from adipose tissue. Thus, excess acetyl-CoA is either metabolized in futile cycles or converted to ketones (β-hydroxybutyrate). Meanwhile, GH and IGF-1 become uncoupled as IGF-1 synthesis is reduced in the liver. The catabolism of muscle produces myostatin, which acts on muscle tissue and adipose tissue disrupting the translocation of Glut 4 to the cell membrane (Antony et al., 2007), resulting in tissues being less responsive to the actions of insulin. To make things worse, tissues become even more unresponsive to insulin when concentrations of ketones and NEFA are in abundance in the peripheral circulation as they too disrupt the translocation of Glut 4 to the cell membrane (Dresner et al., 1999; Schmitz-Peiffer et al., 1999; Tardif et al., 2001). Elevated concentrations of cortisol are also responsible for direct inhibition of glucose transport into cells. As a result, the cascade of events discussed above render it difficult to formulate diets for range beef cows in negative energy balance, and prolonged periods of negative energy balance can deplete maternal energy stores resulting in diminished production (milk, calf gain, resumption of estrus, and delayed conception). In order to regain homeorhesis [i.e., coordination of metabolism in various tissues to support a given physiological state (Bauman and Currie, 1980)], dietary nutrients must be sufficient to redirect nutrient partitioning back toward maternal tissue. Figure 1 illustrates the metabolic signals that are produced and the tissues they affect. As evident by the figure, one can see the complexity of the intertwined signals involved in regulating energy balance in the cow. Mulliniks et al., (2009) summarized 7-years of work conducted at the Corona Range and Livestock Research Center, Corona, New Mexico. These studies evaluated postpartum supplementation strategies in young beef cows with various concentrations of glucogenic precursors [i.e., Ca-propionate and ruminally undegradable protein (containing glucogenic AA)]. Range supplements with greater concentrations of glucogenic precursors decreased days to first estrus and improved pregnancy rates in 2- and 3-yr-old cows. This was accomplished by repartitioning nutrients away from lactation and toward maternal tissues. Furthermore, the addition of Ca-propionate in range supplements fed to cows improved pregnancy rates. The higher cost of adding Ca-propionate to supplements was offset by increasing subsequent year calf crop, which also increased ranch revenue compared with feeding other supplements. Seasonal Changes during the Production Cycle It is rare for dairy cows to experience negative energy balance at times other than just prior to parturition and early in lactation. This is a result of total mixed rations and the quality of feedstuffs used in diets to meet the requirements of dairy cows. On the other hand, range beef cows are extremely dependent upon environmental conditions and the response of rangeland forages to these changing conditions. It is not improbable for range beef cows to experience 2 or 3 periods of weight loss throughout a production cycle in addition to that which occurs around 97 strategically implementing supplementation protocols that target periods of nutritional stress. Identification of these periods of time throughout a production cycle will encourage better utilization of low quality forages, improve nutrient utilization, and lend to achieving production goals for livestock grazing western rangelands. These supplementation strategies may include separation of cattle based on their nutrient status. Nutritionally stressed cows may be eligible to receive supplements ranging from a small packaged, biologically potent, supplement incorporated in a loose mineral mix or delivery of up to 0.9 kg/d of a 40% protein supplement (Petersen, 2006). The key is early intervention, which may result in supplementing livestock during times of the year that traditionally may not have been done or no intervention at all. Understanding the consequences to livestock experiencing a state of negative energy balance are the keys to creating more efficient strategies. FUTURE STRATEGIES Range beef cows are reliant upon the quality and quantity, which are controlled by environmental conditions, of rangeland forages to maintain positive energy balance. Range beef cows in arid and semi-arid environments may only have adequate supplies of nutrients to maintain positive energy balance 40% of the time. Thus, it is essential that strategic supplementation regimes be used in a manner that is economically and biologically attainable. Researchers at Fort Keogh, Livestock and Range Research Laboratory are striving to identify the onset of negative energy balance in beef cows and use biologically potent strategic supplementation to correct this energy imbalance. Glucose tolerance tests are routinely used to measure glucose half-life, glucose area under the curve, and insulin area under the curve as a measure of the onset or severity of negative energy balance. However, these tests are time consuming and the analytic procedures are conducted long after the time when changes could be implemented. We are now evaluating the efficacy of using chute side measures (i.e., as ketone meters), which may allow beef cows to be sorted out of a chute within 10 sec of placing a single drop of blood on the meter’s ketone test strip. This may become a viable tool for producers to make management decisions that allow different feedstuffs to be fed to groups of cows experiencing the same metabolic condition. This may be an approach that strategically allows producers to manage their cows better and be more economically stringent with their supplemental feed. LITERATURE CITED Abdoun, K., F. Stumpff, and H. Martens. 2006. Ammonia and urea transport across the rumen epithelium: a review. Anim. Health Res. Rev. 7:43-59. Adams, D. C. and R. E. Short. 1988. The role of animal nutrition on productivity in a range environment. Pages 37-45 In: Achieving Efficient Use of Rangeland Resources. R. S. White and R. E. Short (eds) Fort Keogh Res. Symp. Miles City, MT. Antony, N., J. J. Bass, C. D. McMahon, and M. D. Mitchell. 2007. Myostatin regulates glucose uptake in BeWo cells. Amer. J. Physiol. Endocrinol. Metab. 293: E1296-1302. Bauman, D. E. and W. B. Currie. 1980. Partitioning of nutrients during pregnancy and lactation: a review of mechanisms involving homeostasis and homeorhesis. J. Dairy Sci. 63:1514-1529. Bell, A. W. and D. E. Bauman. 1997. Adaptations of glucose metabolism during pregnancy and lactation. J. Mammary Gland Biol. Neoplasia 2:265-278. Bergman, E. N. 1973. Glucose metabolism in ruminants as related to hypoglycemia and ketosis. Cornell Vet. 63: 341-382. Bloom, S. R., A. V. Edwards, R. N. Hardy, K. Malinkowska, and M. Slilver. 1975. Cardiovascular and endocrine responses in the young calf. J. Physiol. (Lond) 244:135-155. Bowen, J. M. 1964. Peripheral, hepatic and non-hepatic splanchnic effects of insulin in sheep. Cornell Vet. 54:57-65. Bradford, B. J. and M. S. Allen. 2008. Negative energy balance increases periprandial ghrelin and growth hormone concentrations in lactating dairy cows. Domest. Anim. Endocrinol. 34:196-203. Brockman, R. P. 1978. Roles of glucagon and insulin in the regulation of metabolism in ruminants - A Review. Can. Vet. J. 19:55-62. Butler, W. R. 2000. Nutritional interactions with reproductive performance in dairy cattle. Anim. Reprod. Sci. 60-61:449-457. SUMMARY In extensive, range forage-based production systems, beef cows are expected to adapt to vast array of changing environmental conditions. These changing environmental conditions often lead to unfavorable metabolic states that limit animal performance and subsequently impact production efficiency. Thus, range livestock producers and managers must strive to identify livestock best fitted to produce in their unique environment and management scheme. Unfortunately, none of us have a “Crystal Ball” to tell us when forage quality will diminish in a given year, but we can rely on some historical data to help make management decisions. For the Northern Great Plains at Fort Keogh, 75 to 90% of annual range forge production occurs by July 1, and it is at the 90% forage production level in areas where cool season grasses (C3) are dominate (Heitschmidt and Vermeire, 2005). So why is this important? If an assessment of forage production is made in July a producer can adjust stocking rates to ensure that livestock grazing rangelands have ample opportunity to consume as much forage as they desire (forge intake therefore is not a limiting factor). This leaves the question to quality and whether range livestock can consume sufficient nutrients to achieve a desired production goal. This paper presented the metabolic consequences to range livestock when they are unable to consume sufficient nutrients to attain positive or neutral energy balance. Research underway at Fort Keogh and previously reported from New Mexico State University has focused on 98 Chagas, L. M., J. J. Bass, D. Blache, C. R. Burke, J. K. Kay, D. R. Lindsay, M. C. Lucy, G. B. Martin, S. Meier, F. M. Rhodes, J. R. Roche, W. W. Thatcher, and R. Webb. 2007. Invited review: New perspectives on the roles of nutrition and metabolic priorities in the subfertility of high-producing dairy cows. J. Dairy Sci. 90:4022-4032. Chilliard, Y., A. Ferlay, Y. Faulconnier, M. Bonnet, J. Rouel, and F. O. Bocquier. 2000. Adipose tissue metabolism and its role in adaptations to undernutrition in ruminants. Proc. Nutr. Soc. 59:127-134. Croom, W. J., Jr., L. S. Bull, and I. L. Taylor. 1992. Regulation of pancreatic exocrine secretion in ruminants: A review. J. Nutr. 122:191-202. Dresner, A., D. Laurent, M. Marcucci, M. E. Griffin, S. Dufour, G. W. Cline, L. A. Slezak, D. K. Andersen, R. S. Hundal, D. L. Rothman, K. F. Petersen, and G. I. Shulman. 1999. Effects of free fatty acids on glucose transport and IRS-1-associated phosphatidylinositol 3kinase activity. J. Clin. Invest. 103:253-259. Du, M., J. Tong, J. Zhao, K. R. Underwood, M. Zhu, S. P. Ford, and P. W. Nathanielsz. 2010. Fetal programming of skeletal muscle development in ruminant animals. J. Anim. Sci.88:E51-60. Fahey, G. C., Jr. and L. L. Berger. 1987. Carbohydrate nutrition of ruminants. Pages 269-297 In: The Ruminant Animal: Digestive Physiology and Nutrition. D. C. Church (ed) Simon & Schuster, Englewood Cliffs, NJ. Faulkner, A. and P. A. Martin. 1997. The concentrations of some gut polypeptides are elevated during lactation in ruminants - a potent glycogenic hormone. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 118:563568. Forbes, J. M. 1982. The role of the liver in the control of food intake. Proc. Nutr. Soc. 41:123-126. Friedman, J. M. and J. L. Halaas. 1998. Leptin and the regulation of body weight in mammals. Nature 395:763770. Grings, E. E., R. E. Short, K. D. Klement, T. W. Geary, M. D. MacNeil, M. R. Haferkamp, and R. K. Heitschmidt. 2005. Calving system and weaning age effects on cow and preweaning calf performance in the Northern Great Plains. J. Anim. Sci. 83:2671-2683. Harrelson, F. W., S. L. Lodge-Ivey, S. H. Cox, J. T. Dunlap, J. T. Mulliniks, B. H. Carter, C. A. Loest, and M. K. Petersen. 2009. Effect of winter supplementation strategy on glucose metabolism pre- and postpartum. Proc. West. Sec. Amer. Soc. Anim. 60:354-357. Hawkins, D. E., M. K. Petersen, M. G. Thomas, J. E. Sawyer, and R. C. Waterman. 2000. Can beef heifers and young postpartum cows be physiologically and nutritionally manipulated to optimize reproductive efficiency? J. Anim. Sci. 77:1-ar-10-ar. Hayashida, T., K. Murakami, K. Mogi, M. Nishihara, M. Nakazato, M. S. Mondal, Y. Horii, M. Kojima, K. Kangawa, and N. Murakami. 2001. Ghrelin in domestic animals: distribution in stomach and its possible role. Domest. Anim. Endocrinol. 21:17-24. Heiman, M. L., K. W. Sloop, J. F. Caro, and Y. Chen. 1999. Extension of neuroendocrine axes to include leptin. J. Anim. Sci. 77:33-42. Heitschmidt, R. K., E. E. Grings, M. R. Haferkamp, and M. G. Karl. 1995. Herbage dynamics on two Northern Great Plains range sites. J. Range. Manage. 48:211-217. Heitschmidt, R. K., M. R. Haferkamp, M. G. Karl, and A. L. Hild. 1999. Drought and grazing: I. Effects on quantity of forage produced. J. Range Manage. 52:440446. Heitschmidt, R. K. and L. T. Vermeire. 2005. An ecological and economic risk avoidance drought management decision support system. In: Pastoral Systems in Marginal Environments. J. A. Milne (ed) Proc. Satellite Workshop 20th Interntl. Grassland Congr. July 3-6. Glasgow, Scotland. The Netherlands: Wageningen Academic Publ. 178. Hess, B. W., S. L. Lake, E. J. Scholljegerdes, T. R. Weston, V. Nayigihugu, J. D. C. Molle, and G. E. Moss. 2005. Nutritional controls of beef cow reproduction. J. Anim. Sci. 83:E90-E106. Hulshof, S. C., J. R. Figueiredo, J. F. Beckers, M. M. Bevers, and R. van den Hurk. 1994. Isolation and characterization of preantral follicles from foetal bovine ovaries. Vet. Q. 16:78-80. Jarrett, I. G., O. H. Filsell, and F. J. Ballard. 1974. Metabolic and endocrine interrelationships in normal and diabetic sheep. Horm. Metab. Res. Suppl. 4:111116. Kaneko, J. J. 1989. Clinical Biochemistry of Domestic Animals. 4th ed. Academic Press, San Diego, CA. Keisler, D. H. and M. C. Lucy. 1996. Perception and interpretation of the effects of undernutrition on reproduction. J. Anim Sci.(Suppl. 3) 74:1-17. Khachadurian, A. K., B. Adrouni, and H. Yacoubian. 1966. Metabolism of adipose tissue in the fat tail of the sheep in vivo. J. Lipid Res. 7:427-436. Lee, S.-J. 2004. Regulation of muscle mass by myostatin. Ann. Rev. Cell Dev. Biol. 20:61-86. Lucy, M. C., G. A. Verkerk, B. E. Whyte, K. A. Macdonald, L. Burton, R. T. Cursons, J. R. Roche, and C. W. Holmes. 2009. Somatotropic axis components and nutrient partitioning in genetically diverse dairy cows managed under different feed allowances in a pasture system. J. Dairy Sci. 92:526-539. McDonald, I. W. 1948. The absoption of ammonia from the rumen of the sheep. Biochem. J. 42:584-587. McShane, T. M., S. L. Petersen, S. McCrone, and D. H. Keisler. 1993. Influence of food restriction on neuropeptide-Y, proopiomelanocortin, and luteinizing hormone-releasing hormone gene expression in sheep hypothalami. Biol. Reprod. 49:831-839. Mulliniks, J. T., S. H. Cox, M. E. Kemp, R. L. Endecott, R. C. Waterman, D. M. VanLeeuween, and M. K. Petersen. 2009. Increasing glucogenic precursors in range supplements improves reproductive efficiency and profitability in young postpartum range cows in years 2000-2007. Proc. West. Sec. Amer. Soc. Anim. 60:7680. 99 Parker, D. S., M. A. Lomax, C. J. Seal, and J. C. Wilton. 1995. Metabolic implications of ammonia production in the ruminant. Proc. Nutr. Soc. 54:549-563. Pell, J. M., and E. N. Bergman. 1983. Cerebral metabolism of amino acids and glucose in fed and fasted sheep. Amer. J. Physiol. Endocrinol. Metab. 244:E282-E289. Petersen, M. K. 2006. Strategic low cost supplementation. Proc. West. Sec. Amer. Soc. Anim. 57:3-6. Prasad, B. M., C. D. Conover, D. K. Sarkar, J. Rabii, and J. P. Advis. 1993. Feed restriction in prepubertal lambs: effect on puberty onset and on in vivo release of luteinizing-hormone-releasing hormone, neuropeptide Y and beta-endorphin from the posterior-lateral median eminence. Neuroendocrinol. 57:1171-1181. Relling, A. E. and C. K. Reynolds. 2007. Plasma Concentrations of Gut Peptides in Dairy Cattle Increase after Calving. J. Dairy Sci. 90:325-330. Reynolds, C. K. 1992. Metabolism of nitrogenous compounds by ruminant liver. J. Nutr. 122:850-854. Roche, J. R., D. Blache, J. K. Kay, D. R. Miller, A. J. Sheahan, and D. W. Miller. 2008. Neuroendocrine and physiological regulation of intake with particular reference to domesticated ruminant animals. Nutr. Res. Rev. 21:207-234. Salway, J. G. 1994. Metabolism At A Glance. Blackwell Science Ltd., Cambridge, MA. Schmitz-Peiffer, C., D. L. Craig, and T. J. Biden. 1999. Ceramide generation is sufficient to account for the inhibition of the insulin-stimulated PKB pathway in C2C12 skeletal muscle cells pretreated with palmitate. J. Biol. Chem. 274:24202-24210. Short, R. E., R. A. Bellows, R. B. Staigmiller, J. G. Berardinelli, and E. E. Custer. 1990. Physiological mechanisms controlling anestrus and infertility in postpartum beef cattle. J. Anim. Sci. 68:799-816. Sugino, T., J. Yamaura, M. Yamagishi, A. Ogura, R. Hayashi, Y. Kurose, M. Kojima, K. Kangawa, Y. Hasegawa, and Y. Terashima. 2002. A transient surge of ghrelin secretion before feeding is modified by different feeding regimens in sheep. Biochem. Biophys. Res. Commun. 298:785-788. Tardif, A., N. Julien, A. Pelletier, G. Thibault, A. K. Srivastava, J. L. Chiasson, and L. Coderre. 2001. Chronic exposure to beta-hydroxybutyrate impairs insulin action in primary cultures of adult cardiomyocytes. Amer. J. Physiol. Endocrinol. Metab. 281:E1205-E1212. Thomas, M. G., O. S. Gazal, G. L. Williams, R. L. Stanko, and D. H. Keisler. 1999. Injection of neuropeptide Y into the third cerebroventricle differentially influences pituitary secretion of luteinizing hormone and growth hormone in ovariectomized cows. Domest. Anim. Endocrinol. 16:159-169. Tschop, M., D. L. Smiley, and M. L. Heiman. 2000. Ghrelin induces adiposity in rodents. Nature 407:908913. van Knegsel, A. T. M., H. van den Brand, J. Dijkstra, and B. Kemp. 2007. Effects of dietary energy source on energy balance, metabolites and reproduction variables in dairy cows in early lactation. Theriogenol. 68:S274S280. van Knegsel, A. T. M., H. van den Brand, J. Dijkstra, S. Tamminga, and B. Kemp. 2005. Effect of dietary energy source on energy balance, production, metabolic disorders and reproduction in lactating dairy cattle. Reprod. Nutr. Dev. 45:665-688. Waterman, R. C., E. E. Grings, T. W. Geary, A. J. Roberts, L. J. Alexander, and M. D. MacNeil. 2007. Influence of seasonal forage quality on glucose kinetics of young beef cows. J. Anim. Sci. 85:2582-2595. Waterman, R. C., J. E. Sawyer, C. P. Mathis, D. E. Hawkins, G. B. Donart, and M. K. Petersen. 2006. Effects of supplements that contain increasing amounts of metabolizable protein with or without Ca-propionate salt on postpartum interval and nutrient partitioning in young beef cows. J. Anim. Sci. 84:433-446. Wertz-Lutz, A. E., T. J. Knight, R. H. Pritchard, J. A. Daniel, J. A. Clapper, A. J. Smart, A. Trenkle, and D. C. Beitz. 2006. Circulating ghrelin concentrations fluctuate relative to nutritional status and influence feeding behavior in cattle. J. Anim. Sci. 84:3285-3300. West, C. E. and R. F. Passey. 1967. Effects of glucose load and of inslulin on the metabolism of glucose and of palmitate in sheep. Biochem. J. 102:58-64. Zimmers, T. A., M. V. Davies, L. G. Koniaris, P. Haynes, A. F. Esquela, K. N. Tomkinson, A. C. McPherron, N. M. Wolfman, and S.-J. Lee. 2002. Induction of cachexia in mice by systemically administered myostatin. Science 296:1486-1488. 100 101 Figure 1. Metabolic signals (releasing and target tissues) that are influenced by energy balance of beef cows (POA = preoptic area of the hypothalamus, MBH = medial basal hypothalamus, AP = anterior pituitary, and PP = posterior pituitary). METABOLIC SIGNALS OF THE BEEF COW IN NEGATIVE ENERGY BALANCE1 R. C. Waterman and W. R. Butler Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 MATERNAL PLANE OF NUTRITION: IMPACTS ON FETAL OUTCOMES AND POSTNATAL OFFSPRING RESPONSES1,2 J. S. Caton3ŧ and B. W. Hess* ŧ Center for Nutrition and Pregnancy, Animal Sciences Department, North Dakota State University, Fargo 58108 *Department of Animal Science, University of Wyoming, Laramie 82071 ABSTRACT: Objectives of this review are to summarize the current information regarding effects of maternal nutrition on fetal outcomes and postnatal offspring performance and provide a platform to assist in the formation of additional hypotheses regarding relevant aspects of developmental programming in ruminant livestock species. Inadequacies in maternal nutrition have been implicated in developmental programming and resulting pre and postnatal changes that affect long-term offspring halth and performance. Developmental programming is the concept that perturbations during critical developmental periods may have long-term impacts on offspring outcomes. Although this concept not new, both early epidemiological studies and subsequent follow up work with animal models has increased interest among many biological scientists. Within extensive ruminant livestock production systems, periods of nutrient inadequacy are common and present production challenges. Decreased growth rate and suboptimal carcasses cost feedlot producers millions of dollars annually. In addition, fetal growth restriction resulting from perturbed maternal nutrition has been associated with negative impacts upon developing tissues, metabolism, growth efficiency, and body composition. The degree to which inadequacies in maternal nutrition impacts efficiencies and profitability of ruminant livestock production systems remains to be fully determined. INTRODUCTION Maternal nutritional plane (Barker, 2004; Wu et al., 2006) has been implicated in developmental programming and resulting pre and postnatal changes that affect longterm offspring health and performance. Developmental programming is the concept that perturbations during critical prenatal or postnatal developmental stages can have lasting impacts on growth and adult function. Maternal nutritional status (Wallace, 1948; Wallace et al., 1999; Godfrey and Barker, 2000; Wu et al., 2006) is a major factor implicated in development and function of the fetal organ systems. Effective nutritional (Bull and Carroll, 1937; Barcroft, 1946; Morrison, 1949; Maynard and Loosli, 1956; Crampton and Lloyd, 1959; NRC, 1970) management during gestation has long been recognized as an important component of sound livestock production practices. Fetal intrauterine growth restriction (IUGR) is an issue in relevant livestock species in certain circumstances (Wu et al., 2006) and can be precipitated by numerous forms of perturbed maternal nutrition. Recent observations and reviews in the area of developmental programming (Barker et al., 1992; Godfrey and Barker, 2000; Armitage et al., 2004; Barker, 2004; Wu et al., 2006) have elevated the concept of maternal onset of adult disease and lifelong performance to the forefront of investigation in numerous national and international laboratories. In addition, underlying concepts of developmental programming are driving the development of research programs in multiple laboratories and shaping thought in funding arenas in both biomedical and animal agriculture. This review will focus on the impacts of maternal nutritional plane on resulting fetal and postnatal outcomes in relevant livestock species. Emphasis will also be placed on growth, development, metabolic, and performance outcomes, timing of nutritional perturbations, potential long-term production consequences, and areas for future research. Key words: developmental programming, fetal growth, maternal nutrition, offspring performance, ruminants 1 Mention of trade names or commercial products in this publication is solely for the purpose of providing specific information and does not imply recommendation or endorsement by the authors. 2 The authors would like to thank the numerous collaborators that have contributed to the data and concepts contained within this manuscript. Appreciation is also expressed to the organizing committee of the Grazing Nutrition Conference. This project was partially supported by the National Research Initiative Competitive Grants no. 2003-35206-12814, 2005-35206-15281, 2006-5561816914, and 2009-35206-05253 from the USDA Cooperative State Research, Education, and Extension Service. 3 Corresponding author: [email protected] MATERNAL NUTRITIONAL PLANE AND DEVELOPMENTAL PROGRAMMING Developmental Programming Developmental programming, also termed “Fetal Programming,” “The Barker Hypothesis,” or, “developmental origins of health and disease,” is a major concept underpinning research in many biomedical and agricultural laboratories (Hanson and Gluckman, 2005; Wu 104 confer an immediate survival advantage to the offspring assuming the experienced prenatal environment is indicative of the ensuing postnatal environment. An example would be in the case of maternal nutrient restriction resulting in developmental changes that predispose the offspring for better survival in a nutrient limiting environment. However, if the postnatal environment does not match the prenatal environment the offspring may have adverse consequences that resemble aspects of metabolic syndrome later in life. Aspects of developmental programming have been described by both the thrifty phenotype hypothesis and by predictive adaptive responses. In addition severe perturbations during development can have dramatic impacts on developmental and offspring outcomes. Certainly, severe malnutrition has both immediate and long term consequences to developing offspring, as would infectious agents or toxins. These adverse affects during development from sever circumstances are more pronounced and often not viewed as adaptive responses, but more as consequences of extreme perturbations (Hanson and Gluckman, 2005). However, the importance of such examples should not be overlooked in providing insight into mechanism driving more modest responses to less compromised intrauterine environments. Clearly, maternal nutritional status is one of the factors implicated in programming, nutrient partitioning, and ultimately growth, development, and function of the major fetal organ systems (Wallace, 1948; Wallace et al., 1999; Godfrey and Barker, 2000; Wu et al., 2006; Caton et al., 2007). The prenatal growth trajectory is sensitive to direct and indirect effects of maternal dietary intake from the earliest stages of embryonic life when the nutrient requirements for conceptus growth are negligible (Robinson et al., 1999). This is especially relevant because pre-term delivery and fetal growth restriction are associated with greater risk of neonatal mortality and morbidity. Growth restricted neonates are not only at risk of immediate postnatal complications, but may also exhibit poor growth and development, with significant consequences later in life (Barker et al., 1993; Godfrey and Barker, 2000; Barker, 2004; Wu, 2006). Because of accumulating data and potential ramifications, the concept that a maternal stimulus, including plane of nutrition may result in developmental programming responses that impact neonatal, growth and development, and adult responses is gaining traction in animal agriculture. et al., 2006). Developmental programming is the concept that perturbations during critical developmental periods may have long-term “programmed” impacts on offspring outcomes. While not new, this concept has piqued interest among biological scientists by both early human epidemiological studies (Barker et al., 1989; Barker, 1994, 2004) and subsequent follow up work with animal models. Barker and his colleagues evaluated human birth record data, and related maternal nutrient restriction and infant weight and physical characteristics at birth to health status in later life. They found standard mortality ratios in men and in women decline as birth weight increased. They further indicated that low birth weight infants tend to have higher mortality ratios from coronary heart disease. Based upon these data, they suggest that maternal nutrition restriction during the first half of pregnancy can lead to smaller for gestational age infants at birth resulting in increased likelihood of health problems, including obesity, diabetes, and cardiovascular disease, that can be experienced later in life. Later work has suggested that similar responses could also be observed in the absence of reduced birth weight. Building upon these concepts (Hales and Barker, 1992) proposed the “Thrifty Phenotype Hypothesis” which was originally put forward to explain the relationship between low birth weight and type-2 diabetes. The thrifty phenotype hypothesis indicates that offspring born from nutrient restricted or malnourished intrauterine environments will have a competitive advantage if they are born into postnatal environments that are also nutrient limiting. Later, Hales and Barker (2001) expanded this hypothesis to explain fetal origins of metabolic syndrome (sometimes called syndrome X). In humans, metabolic syndrome is defined as possessing three of the following five conditions: high blood pressure, elevated abdominal adiposity, increased serum triglycerides, decreased serum HDL cholesterol, and fasting hyperglycemia (Wilson and Grundy, 2003 a,b). Others have suggested additional conditions for inclusion into metabolic syndrome (Armitage et al., 2004) including endothelial dysfunction, which some (Pinkney et al., 1997; Bonora et al., 2003) have suggested as a potential underlying cause. Naturally, these issues are of paramount concern in human biomedical areas, particularly in light of increasing obesity and metabolic syndrome in populations of developed countries. From a livestock production perspective, both internal and external fat are economically important and related to carcass grading and reproductive efficiency. Proper vascular, and by association, endothelial functions also are critical for nutrient exchange in key nutrient transferring tissues (intestine, placenta, kidney, and mammary gland) that support livestock production and health. Responses to environmental perturbations including nutrient restriction and maternal stress have been termed “predictive adaptive responses” (Gluckman and Hanson, 2004a,b; Hanson and Gluckman, 2005). This concept differs from the thrifty phenotype hypothesis discussed above in that predictive adaptive responses confer a benefit later in life (Hanson and Gluckman, 2005; Martin-Gronert and Ozanne, 2006). The thrifty phenotype hypothesis attempts to explain intrauterine developmental changes that Practical Relevance in Livestock Within livestock production systems, there is real potential for ruminants to undergo periods of undernutrition (extensive grazing or high milk output) or overnutrition (overfeeding) during gestation. Decreased growth rate and suboptimal carcasses cost feedlot producers millions of dollars annually (Smith et al., 1995; Gardner et al., 1998). Fetal growth restriction and maternal undernutrition are implicated in negative impacts upon growth efficiency and body composition (Greenwood et al., 1998, 2000; Wu et al., 2006; Caton et al., 2007; Larson et al., 2009). Maternal nutrient restriction can significantly alter composition of 105 historical concept is that during pregnancy nutrient partitioning favors the conceptus at the expense of the dam (Barcroft, 1946). However, exposure of pregnant sheep to severe maternal undernutrition at all stages of pregnancy, and particularly during late gestation, reduces fetal growth (Mellor, 1983; Robinson, 1983; Vincent et al., 1985; Parr et al., 1986). Nutrient Restriction. Nutrient restriction is broadly defined as any series of events that reduce fetal and (or) perinatal nutrient supply during critical windows of development. Nutrient restriction can result from altered maternal nutrient supply, placental insufficiency, deranged metabolism and regulation, physiological extremes, and environmental conditions. From a practical standpoint, maternal nutrient supply and environmental conditions leading to stress responses are the most likely observed causes of nutrient restriction in ruminant livestock. However, the incidence of multiple births in some breeds can also contribute to physiological extremes and result in reduced nutrient supply to developing fetuses. We will focus primarily on maternal dietary-induced nutrient restriction in this review. Due to the pattern of placental growth in relation to fetal growth during gestation (Redmer et al., 2004), it is important to realize that the effects of nutrient restriction during pregnancy may depend on the timing, level, and (or) length of nutrient restriction. In an excellent review, Luther et al., (2005) indicated that maternal nutrient restriction in sheep during and the through mid pregnancy could reduce placenta size and function, while having minimal impacts on fetal body weight near term. Additionally, Reed et al. (2007) and Swanson et al. (2008) demonstrated that maternal nutrient restriction during the last two thirds of pregnancy in sheep reduces birth weights. Sletmoen-Olson et al. (2000) indicated that both low and high levels of metabolizable protein supplementation to mature beef cows reduce birth weights relative to controls fed at projected requirement. In contrast, protein supplementation of cows during the last trimester of pregnancy has been reported to have little effect on birth weights (Martin et al., 2007; Larson et al., 2009). In addition, a moderate level of total nutrient restriction during the second (Lake et al., 2005) and last two thirds of gestation (H. Freetly, personal communication) in mature beef cows has not impacted birth weights. From the available data it appears that birth weights in sheep are more susceptible to maternal nutrient restriction than beef cattle. Nutrient Excess. Increasing plane of nutrition to help beef cows achieve a body condition score of 6 (9-point scale) at parturition did not affect calf birth weights (Lake et al., 2005), but others (Spitzer et al., 1995 and Stalker et al., 2006) reported that increasing body condition before parturition can increase calf birth weight. Maternal overnutrition during gestation can also have adverse effects. It has been shown that overnourishing the singleton-bearing adolescent ewe throughout gestation results in rapid maternal growth, and most particularly of maternal adipose tissue, at the expense of the nutrient requirements of the gravid uterus (Wallace et al., 1996, 1999, 2001). In this paradigm, rapid maternal growth results in placental growth restriction, premature delivery of low-birth weight lambs offspring growth in the absence of birth weight differences (Gardner et al., 2005; Ford et al., 2007). Permanent changes in postnatal metabolism induced by maternal nutritional perturbations may present significant challenge to livestock producers because nutritional management decisions are often based on, for example, average body weight of a given group of animals, which may result in undernutrition of animals of below average body weight or overnutrition of animals heavier than average. Therefore, information regarding factors contributing to animal variation, such as developmental programming, has potential to improve efficiency and profitability of rearing programs. Investigating maternal nutritional effects on offspring growth is very economically relevant to agricultural producers. Weaning weight is an important factor affecting profitability for producers who sell their calves at weaning (Reed et al., 2006). Fetal undernutrition frequently occurs in animal agriculture, leading to reduced fetal growth (Wu et al., 2006). Nutritional requirements of spring-calving beef cows grazing dormant range during late gestation typically exceed the nutrient value of the grazed forage (NRC, 1996; Johnson et al., 1998; Cline et al., 2009: Lardy and Caton, 2010). Gestating ewes grazing rangeland pastures often experience prolonged bouts of undernutrition (Thomas and Kott, 1995) and the nutrient uptake of grazing ewes in the western United States is often less than 50% of NRC recommendations (NRC, 1985). Maternal nutrition pre- and postpartum has been shown to affect calf growth through weaning (Perry et al., 1991; Beaty et al., 1994; Spitzer et al., 1995; Stalker et al., 2006), with greater weaning weights produced from cows on a greater plane of nutrition. In most of the previously mentioned studies offspring were allowed to nurse dams until weaning. The true effect of maternal nutrition on offspring is difficult to decipher when offspring are left nursing their dams until weaning, since prepartum nutrition and postpartum nutrient supply (lactation), both contribute to postnatal growth and performance. Some investigators (Greenwood et al., 1998, 2000; Caton et al., 2007; Swanson et al., 2008; Neville et al., 2010; Meyer et al., 2010b) have raised offspring independent of postnatal maternal influences, which should eliminate the confounding effects of prenatal and postnatal nutritional supply on offspring performance. Maternal Nutrition Maternal nutrient supply in terms of total nutrient intake or specific nutrient availably can vary widely in practical conditions and can be driven by dietary quantity and (or) quality issues, perturbations of intake, and environmental conditions. Maternal nutritional status is one of the factors implicated in programming postnatal nutrient partitioning and ultimately growth and development, including that of muscle and adipose tissue (Wallace, 1948; Godfrey and Barker, 2000; Armitage et al., 2005; Wu et al., 2006; Taylor and Poston, 2007). Growth restricted infants are not only at risk of immediate postnatal complications, but also may be „programmed‟ to exhibit poor growth and productivity and also to develop significant diseases later in life (Barker et al., 1993; Godfrey and Barker, 2000; Wu et al., 2006). The 106 compared with moderately nourished ewes of equivalent age. Swanson et al. (2008) also demonstrated that birth weights are decreased by 9.2% when adolescent ewes are fed 140% of energy requirements from d 40 of pregnancy until parturition, which indicates that moderate overnutrition during the last two-thirds of pregnancy can cause moderate fetal growth restriction. Although placental and fetal growth restriction is generally seen only in adolescent, overnourished ewes, gestation length, and colostrum yield are negatively affected in adult ewes overnourished throughout pregnancy (Wallace et al., 2005). Thus, the health and growth of offspring from adult animals may also be altered by maternal overnutrition. Animal Models and Mechanisms of Developmental Programming Livestock animal models have and will continue to provide significant impacts in critical areas of nutritional and biomedical research. In fact, our current understanding of nutritional principles and applications has been greatly impacted by both discovery and application-based research with livestock (Reynolds et al., 2009; 2010b). Briefly, the broad nutritional areas of energetics (Nichols and Reeds, 1991; Johnson, 2007), proteins (Mathewes, 1991; Baracos, 2004; Bergen, 2007), carbohydrates and lipids (Bergen and Mersmann, 2005; Nafikov and Beitz, 2007), vitamins (McDowell, 2000), minerals (O‟Dell and Sunde, 1997; Underwood and Suttle, 1999; Sandstead and Klevay, 2000; McDowell, 2003), and growth and body composition (Mitchell, 2007) have been largely impacted by research in livestock species. Likewise, in the case of most known macro and micronutrients, contributions from livestock animal models to our current understanding are extensive. From a historical perspective, nutritional research has progressed as complementary research efforts in agriculturally relevant species, rodents, and humans, with breakthroughs in each fostering complimentary work in the other species (McCollum, 1957; Carpenter, 2003a,b,c,d; Stahl et al., 2008). Consequently, early work in the area of energetics and micronutrients were instrumental in development of the disciplines of biochemistry and subsequently molecular biology (Nichols and Reeds, 1991). Use of appropriate animal models, especially those with agricultural relevance, is and will continue to be critical in nutritional and biomedical research (Baker, 2008; Stahl et al., 2008). Emerging functional examples of continued importance of livestock animal models to a broader understanding of nutrition and biology include, but are not limited to porcine models of obesity and metabolic syndrome (Spurlock and Gabler, 2008), bovine models of fat synthesis regulation (Bauman et al., 2008), and ovine models of developmental programming and pregnancy outcomes (Wallace et al., 2005; Symonds et al., 2007; Barry and Anthony, 2008; Reynolds et al., 2010a). FETAL GROWTH AND CRITICAL WINDOWS: OPPORTUNITIES FOR DEVELOPMENTAL PROGRAMMING Nutrient Supply during Fetal Development The developmental programming concept makes biological sense considering an individual passes more milestones in development in utero than postnatal. The first half of gestation is the time when maximal placental growth, development, and vascularization are occurring. Progressively, as gestation advances blood flow increases to the gravid uterus, and more specifically, nutrient and oxygen exchange at the maternal:fetal interface is vital for optimal fetal growth and development. The ruminant placenta appears unable to extract additional amounts of nutrients per unit of maternal blood in response to fetal demand; therefore, increasing the number of blood vessels at the site of the maternal:fetal exchange is absolutely necessary to optimize blood flow and thus nutrient transfer. Patterns of Fetal Growth in Ruminant Livestock Most growth of the bovine (and other ruminant) fetus occurs during the final third of gestation (Ferrell et al., 1976; Prior and Laster, 1979). Fetal nutrient uptake becomes quantatively important as a contributor to maternal nutrient requirements mid-gestation (Ferrell et al., 1983). Thus, severe nutritional restriction for at least the last half to one-third of pregnancy is usually required to reduce bovine fetal growth. Although placental characteristics may be altered by nutrition during early and mid pregnancy without significantly affecting fetal size (Rasby et al., 1990) or birth weight (Perry et al., 1999; 2002), development and growth of vital organs precedes development of bone, muscle, and fat (Figure 1). Hence, the mass of the relatively late maturing carcass tissues are generally considered more susceptible to the effects of maternal nutrition during later pregnancy when the impacts on fetal growth are likely to be greatest. More subtle effects on organ and tissue development due to nutrition during early pregnancy may occur, with potential for long-term consequences for health (Wu et al., 2006). Critical Developmental Windows The first of 10 principles of developmental programming outlined by Nathanielsz (2006) is that there are critical periods of vulnerability to suboptimal conditions during development. Periods of vulnerability or critical windows differ for specific tissues and organ systems dependent upon developmental stages and timing. During these critical windows, specific challenges like maternal nutritional imbalances can result in changes that have both short and long-term consequences. These include the periconceptual period, placental development and organogensis, rapid fetal growth, and the perinatal period (Figure 2; Fowden et al., 2006; Nathanielsz, 2006; Symonds et al., 2007). In several mammalian species, including livestock, evidence exists that perturbations during these critical windows can alter development and function. Nutrient restriction during the periconceptional period results in impaired blastocyst formation (Borowczyk 107 which maintained body weight and body condition through an 8 week pre-superovulation period, an overfed group that increased adiposity, and a restricted group that lost body weight. The overfed group was fed for ad libitum feed consumption and the restricted group was fed at 60% of controls. After 8 weeks ewes were superovulated and oocytes collected. There was no difference in the number of characterized healthy oocytes in control, overfed, and restricted ewes. However, both the overfed and the restricted ewes had reduced successful fertilizations, morulas, and blastocysts, indicating that maternal nutrition (both inadequate and excess) before mating can have profound impacts on oocyte quality and fertilization rates. et al., 2006), shortened gestation, and abnormal function of the hypothalamic-pituitary axis (Kumarassamy et al., 2005) in sheep. Likewise, insults during various critical stages (Figure 2) can also have measurable and lasting effects on particular tissues or organs undergoing rapid growth and (or) differentiation (see below). PERTURBATIONS OF FETAL DEVELOPMENT AND OFFSPRING RESPONSES IN RUMINANT LIVESTOCK Periconceptional Nutrition and Oocyte Quality Nutritional status is a major factor influencing an animal‟s ability to reproduce (Robinson, 1990; Webb et al., 1999; O‟Callaghan et al., 2000; Hess et al., 2005). Nutrition has a significant impact on numerous reproductive functions including hormone production, fertilization, and early embryonic development (Boland et al., 2001; Armstrong et al., 2003; Boland and Lonergan, 2005). Changing nutrient supply during the peri-conceptual period can have profound impacts on reproduction. Nutritional status has been correlated with embryo survival and is a key factor influencing efficiency in assisted reproductive technologies (Armstrong et al., 2003; Webb et al., 2004). Conflicting results have been reported for the effects of low or high energy diets on oocyte quality and early embryonic development in ruminants (Kendrick et al., 1999; Boland et al., 2001; Papadopoulos et al., 2001). For example, sheep fed low energy diets have embryos with decreased cleavage rates compared with high-energy diets (Papadopoulos et al., 2001). In contrast, a higher proportion of ova from ewes fed low-energy diets were considered viable compared with those from ewes fed high-energy diets (McEvoy et al., 1995). Others (Bloomfield et al., 2003; Kumarasamy et al., 2005) have reported that approximately half of ewes experiencing nutrient restriction during the periconceptual period have marked reductions in gestation length. For cows, positive (Nolan et al., 1998; Kendrick et al., 1999; Boland et al., 2001), negative (Yaakub et al., 1999; Armstrong et al., 2001) or no effects (Tripp et al., 2000) of plane of nutrition (high vs. low energy diets) on oocyte quality, fertilization rate, and early embryonic development have been reported. Recent studies (Borowczyk et al., 2006; Grazul-Bilska et al., 2006) have investigated the effects of maternal plane of nutrition before superovulation and in vitro fertilization on oocyte quality and embryonic development. Borowczyk et al. (2006) compared control ewes fed a maintenance diet to an underfed group that was fed at 60% of controls. After 8 weeks ewes were superovulated, oocytes collected and evaluated, in vitro fertilization rates of oocytes determined, and viability to morula and blastocysts states of embryos determined. There were no differences in the number of healthy oocytes collected from control and restricted ewes. However, at fertilization, restricted ewes produced oocytes that fertilized more poorly. In addition, morulas and blastocysts were less in the underfed group. In a follow-up study, Grazul-Bilska et al. (2006) added an additional over fed treatment. In the second study, which was also with ewes, these researchers had a control group, Placental Development A primary role of the placenta is to provide for physiological interface, including nutrient and waste exchange between the fetal and maternal systems (Meschia, 1983; Reynolds and Redmer, 1995; Reynolds et al., 2010a). Therefore, adequate placental circulation is imperative to successful pregnancy and is exemplified by observed close relationships among fetal weight, placental size, and uterine and umbilical blood flows during normal pregnancies (Reynolds et al., 2005a, 2005b, 2006; 2010a). In the ewe, cotyledonary growth is exponential during the first 10 to 11 wk of pregnancy, with significant slowing until term (Stegeman, 1974: Ferrell et al., 1976). In the cow, the cotyledonary growth progressively increases throughout gestation (Reynolds et al., 1990; Vonnahme et al., 2007). Uterine and umbilical blood flows, which represent the circulation to maternal and fetal portions of the placenta, respectively (Ramsey, 1982; Mossmann, 1987), increase exponentially throughout gestation, essentially keeping pace with fetal growth (Reynolds and Redmer, 1995; Magness, 1998). For example, in sheep the absolute rate of uterine blood flow increases by approximately 3-fold throughout the last half of pregnancy (Meschia, 1983). Over a similar interval of gestation, uterine blood flow in cows increase by 4.5-fold (Reynolds, 1986). As summarized in Table 1, in sheep studied during late gestation, uterine or umbilical blood flows, or both, are reduced in every model of compromised pregnancy in which they have been evaluated. These models of compromised pregnancy include overfed adolescents, underfed adolescent and adult dams, as well as environmental heat-stress, hypoxic stress, and multiple fetuses. In addition, although vascular development of the placenta also is decreased in several of the models of compromised pregnancy, it is increased in others (Table 1). Interestingly, the models in which placental vascularity is increased show no effect on fetal size (high dietary Se or hypoxic stress) or were associated with long-term genetic selection (Romanov vs. Columbia genotype), suggesting an adaptive placental response that preserves the fetal nutrient supply. 108 low-energy diets had fewer fast glycolytic fibers than the offspring of sows fed a high-energy diet during the first 50 d of gestation. These results correspond with results of Schantz et al. (1983), where energy-deficient animals had less fast glycolytic fibers but retained slow oxidative fiber numbers. Muscle fiber type and size not only affect prenatal and postnatal growth, they also can have an impact on endproduct quality. Percentage of slow muscle fibers present can affect meat color (Monin and Ouali, 1992) and speed of tenderization that occurs postmortem (Ouali, 1990). Fiber diameter has an impact on meat tenderness (Maltin et al., 1997). Altering fiber type will also impact glycogen storage capabilities, which has an impact on water-holding capacity (Fernandez and Thornberg, 1991). These product quality traits have a large impact on the acceptability and consumer demand of animal food products. Fetal Organ Development and Postnatal Function Nutrient restriction to the developing conceptus, regardless of the reason (maternal nutrient restriction, environmental conditions, carunclectomy or other experimental models), often results in impaired fetal organogenesis and (or) development. The degree of compromised internal organ growth is usually more severe with increasing extremes of nutrient restriction. Timing of nutrient deprivation can also result in differential effects of fetal organ systems, with restriction during early to mid pregnancy, a critical time for fetal organ growth and development (Figure 1), being more likely to compromise function when compared with moderate levels of restriction during late pregnancy. Various organ systems may respond differently to specific timing and severity of nutrient restriction because of differing growth trajectories and maturation time points. In fact, recent data indicate that both low and high planes of maternal nutrition can impact growth of numerous fetal organs (Reed et al., 2007; Carlson et al., 2009; Caton et al., 2009). Our laboratories we have a particular interest in intestinal development, growth, and function. Intestinal tissues are important to livestock production because of their role in nutrient uptake, immunocompetence, and their disproportional use of energy (and other nutrient resources) in relation to their contribution to overall body mass. A detailed discussion of fetal and offspring intestinal tissue response to perturbations during gestation are beyond the scope of this effort; readers are referred to Reed et al. (2007), Carlson et al. (2009), Caton et al. (2009), Neville et al. (2010), Meyer et al. (2010a), and Zhu et al. (2009). Adipose Tissue Development and Postnatal Accumulation In sheep and cattle, adipose tissue development begins during gestation, but the vast majority of adipose tissue growth occurs postnatally (Mersmann and Smith, 2005). Perirenal adipose tissue represents the primary adipose tissue depot at birth in sheep (Muhlhausler et al., 2002). The perirenal adipose depot undergoes rapid and considerable expansion in the first month of life, during which time the adipose tissue transitions from having a thermoregulatory function to one of energy storage (Clarke et al., 1997). Adipose tissue development during the neonatal period is accompanied by a well-characterized pattern of gene expression as adipose tissue expresses the proteins required for the synthesis of triacylglycerol from glucose and fatty acids (Mersmann and Smith, 2005). Early postnatal growth of adipose tissue is driven by hyperplasia (proliferation) of undifferentiated preadipocytes and subsequent hypertrophy of adipocytes, characterized by the accumulation of triacylglycerol. In adults, adipose tissue growth is driven primarily by hypertrophy of the existing adipocytes, but preadipocyte hyperplasia continues at a low rate for replacement of adipocytes that undergo apoptosis (Mersmann and Smith, 2005). However, marked increases in preadipocyte hyperplasia can occur as a result of excessive caloric intake relative to nutrient requirements (Miller et al., 1984; Shillabeer and Lau, 1994). Clearly, an important concept regarding adipose tissue growth is that animals and humans retain the ability to recruit preadipocytes throughout life, indicating that growth of adipose tissue is essentially unlimited (Hausman et al., 2001). As the lamb grows and approaches maturity, visceral adipose (perirenal and omental tissues) represents approximately one-third of the total adipose tissue mass, with the remainder comprising subcutaneous and intramuscular fat (Moloney et al., 2002). Accumulation of visceral adipose tissue represents a significant inefficiency in meat animal production, considering that much of the visceral fat is discarded at slaughter. While regulation of lipogenesis has depot-specific characteristics as well, relatively few studies have addressed the biological mechanisms by which maternal nutrition alters adipose tissue growth and development in offspring. Additionally, Muscle Development and Product Quality Other more traditionally thought of as “production oriented” tissues like muscle and adipose also appear responsive to programming effects in utero. Maternal nutritional status is one of the factors impacting nutrient partitioning and ultimately growth and development of fetal skeletal muscle (Wallace, 1948; Wallace et al., 1999; Godfrey and Barker, 2000; Rehfeldt at al., 2004; Strickland et al., 2004; Rehfeldt and Kuhn, 2006). Skeletal muscle has a lower priority in nutrient partitioning compared with the brain and heart during development, rendering it potentially more vulnerable to nutrient perturbed nutrient supply (Bauman et al., 1982; Close & Pettigrew, 1990). The fetal period is crucial for skeletal muscle development because no net increase in the number of muscle fibers occurs after birth (Glore & Layman, 1983; Greenwood et al., 2000; Nissen et al., 2003). Numerous studies carried out in a range of mammalian species have shown that maternal undernutrition during gestation can significantly reduce the number of both muscle fibers and nuclei in the offspring (Bedi et al, 1982; Ward and Strickland, 1991). Zhu et al. (2004) observed that nutrient restriction from early to midgestation resulted in a reduction in the number of fetal skeletal muscle fibers, which might be related to a downregulation of mammalian target of rapamycin (mTOR) signaling. Bee (2004) also observed differences in fiber type in semitendinosus muscle where progeny from the 109 treatment, maternal nutritional plane during gestation continued to affect lamb growth through d 19 of age (Meyer et al., 2010b). Specifically, lambs born to ewes restricted to 60% of control intake weighed less than controls at multiple time points up to necropsy at 19 d. Also, lamb ADG decreased in lambs from restricted ewes compared with lambs from control and over fed ewes from birth to d 19. In agreement with the current study, Neville et al., (2010; Table 2) recently reported that lambs born to restricted ewes weighed less at birth than lambs born to control or over-fed dams with some differences persisting to both weaning and 180 d of age. Consistent with these observations, effects of maternal nutrition during gestation on postnatal growth has been reported in cattle (Stalker et al., 2006; Martin et al., 2007; Larson et al., 2009). Growth, digestibility, and N retention were studied in offspring from dams receiving different levels of nutrition and Se during gestation (Neville et al., 2010). As in the work of Meyer et al. (2010b) mentioned above, lambs in these studies were removed from ewes at birth and raised independently. Neville et al. (2010) reported that male offspring from ewes not supplemented with Se had greater ADG then those receiving high Se during gestation. Conversely, male offspring from over fed ewes not supplemented with Se had lower ADG compared with those receiving high Se during gestation. No differences in DMI resulted in the similar responses in Gain:Feed as were found in ADG. Those authors also reported reductions in diet digestibility in growing lambs from ewes fed high Se during gestation when compared with those born to ewes fed adequate Se. No differences were observed in N retention. It was speculated that lambs from high Se dams either had increased passage rates or decreased digestive efficiencies. Others (Martin et al., 2007) found greater residual feed intake in female calves from dams that received a protein supplement during late gestation and an early lactation hay diet. Additionally, Stalker et al. (2006) and Larson et al. (2009) observed no differences in ADG, DMI, or Gain:Feed in steer calves during the finishing period when dams received protein supplements during late gestation. little data are available that addresses the long-term consequences of altered adipose tissue development on body composition and carcass value. The role of maternal nutrition in programming postnatal body composition and growth trajectory may involve permanent alterations in adipocyte morphology, cellularity, and (or metabolism (Taylor and Poston, 2007). In an evaluation of the effects of maternal protein restriction, Guan et al. (2005) revealed marked changes in the characteristics of visceral adipose tissue in the offspring, such as upregulation of genes associated with preadipocyte proliferation, adipocyte differentiation, lipogenesis, and angiogenesis, and a concomitant downregulation of antiangiogenic factors. These data in rodents suggest that, in meat animals, maternal nutrition may alter the trajectory of visceral adipose tissue growth of offspring and prove detrimental to the value of the carcass at slaughter. In sheep, the effects of maternal undernutrition on birth weight and fetal adipose tissue mass have been inconsistent and depend on the timing, level, and (or) length of dietary restriction (Bispham et al., 2003; Symonds et al., 2004). A more consistent observation in lambs from undernourished ewes is that the characteristics of fetal adipose tissue are altered. Other studies have documented increased total visceral (Gardner et al., 2005) and perirenal (Ford et al., 2007) adiposity of the offspring by maternal undernutrition, which was accompanied by insulin resistance compared with offspring from adequately-nourished ewes (Gardner et al., 2005; Gnanalingham et al., 2005; Ford et al., 2007; Caton et al., 2007). Ford et al. (2007) reported that expanded perirenal fat mass in offspring from nutrient-restricted ewes is associated with a reduction in semitendinosus and longissimus muscle mass as a proportion of hot carcass weight. These data, coupled with relative insulin resistance in the offspring from nutrient-restricted dams (Gardner et al., 2005; Ford et al., 2007), indicate that visceral adipose tissue may contribute to abnormal nutrient metabolism. Postnatal Performance Measurements Growth (an increase in the number and size of cells or in mass of tissue) and development (changes in the structure and function of cells or tissues) of the calf are complex biological events. There are numerous growth and development pathways that may influence productive characteristics of ruminants. Growth and development of the bovine fetus, which can be influenced drastically by restricted nutrition during the latter half of gestation (Dunn, 1980) has documented consequences on survival (Holland and Odde, 1992). Likewise, pre-weaning plane of nutrition or level of maternal milk production effects on growth to market weights of cattle are well characterized (Berge, 1991). However, investigations on effects of various treatments or management practices might also need to be made in early- to mid-gestation (Anthony et al., 1986). Long-term consequences of more specific nutritional influences during various stages of fetal and neonatal calf development remain to be determined. When lambs were separated from their dams and raised on an identical management system regardless of maternal DEVELOPMENTAL PROGRAMMING OF OFFSPRING BORN INTO CONVENTIONAL GRAZING LIVESTOCK PRODUCTION SYSTEMS Many extensive beef cattle production systems are designed for cattle to receive the majority of their nutrients from grazed forages. Forage quality is often poor, particularly in dry and winter seasons, and may be inadequate to support optimal nutrition for growth, pregnancy, and lactation without provision of supplemental nutrients. Supplementation of nutrients may not occur for cows during early through mid-gestation because nutrients required to support the growing and developing fetus appear minimal, especially when compared with later stage of gestation and early to mid-lactation. Bovine fetal undernutrition is quite likely to occur during late pregnancy, particularly if cows do not receive supplemental energy and protein. Low precalving plane of nutrition of the cow has been associated with low birth weight of the calf 110 Although fetal weight tended to be reduced, and caruncular (maternal placental) and cotyledonary (fetal placental) weights from nutrient restricted cows were less than control cows on day 125 of gestation, there was no effect of diet on fetal and caruncular weight by day 250 (Zhu et al., 2007). Even after realimentation, cotyledonary weight remained less on day 250 in nutrient restricted cows compared with control cows. Realimentation after ~90 d of nutrient restriction seemed to be the stimulus for altering placental vascularity and development as well as placental function (Vonnahme et al., 2007), resulting in similar fetal weights between nutrient restricted and control cows on day 250 of gestation. It may be misleading to interpret a lack of differences in fetal weights and visceral organ mass of fetuses (Molle et al., 2004), as a result of cow plane of nutrition (Figure 3), to suggest that the offspring will exhibit normal physiological functions later in life. As Figure 4 depicts, both glomerular number and glomeruli proportion of kidney tissue decreased in fetuses gestated by nutrient restricted/realimented cows (Ford et al., 2005). Other researchers (Gilbert et al., 2005) working at the center with lambs born to ewes that were nutrient restricted in early to mid-gestation demonstrated that a 20% reduction in glomerulus number was associated with a 17 mmHg increase in mean arterial pressure (r =–0.85; P <0.01; y =– 5.82x +416). In a companion study with sheep, Vonnahme et al. (2003) reported that both the right and left ventricle of the fetal heart were enlarged if ewes were nutrient restricted during the early to mid-gestation period. At weaning, steer calves from the aforementioned study were placed into a feedlot for 168 days, and pulmonary arterial pressure measurements were taken within a week before steers were slaughtered. Unlike the sheep studies, steers born to nutrient restricted cows did not exhibit an increase in blood pressure (Han et al., 2008). In retrospect, data for steers may be confounded because the bull used for the experiment sired calves which exhibit unusually high pulmonary arterial pressures. Nevertheless, when steers with high pulmonary arterial pressures were compared with steers gestated by nutrient restricted/realimented cow it was noted that 8 common genes were differentially expressed in right ventricles of the heart. Male calves born to cows receiving 68% of their energy requirements from day 31 through 125 of gestation followed by realimentation to achieve similar body condition scores by 250 days of gestation converted feed more efficiently than their control counterparts (Underwood et al., 2008). Greater gain efficiency by steers born to nutrient restricted dams could be related to gain of skeletal muscle because the proportion of lean in 9th through 11th rib section was greater for these animals compared with control steers. Indices of glucose and lipid metabolism did not differ between control steers and steers born to nutrient restricted cows. Although it has not been substantiated, it is interesting to speculate that greater protein gain was related to less protein turnover (Du et al., 2004) and not enhanced protein synthesis (Du et al., 2005). Additional research must be conducted before definitive conclusions can be drawn because preliminary results from recent experiments in our lab would not suggest that feedlot performance and (Bellows et al., 1971), and calf birth weight was increased in heifers and cows in response to dietary supplementation with protein and energy concentrates during late gestation (Clanton and Zimmerman, 1970; Bellows and Short, 1978). Consequences of Poor Plane of Nutrition Early to MidGestation Nutritional inadequacy is likely to occur in beef cattle production systems in which cows are managed to produce a calf annually. In these systems, with an average gestation length of 285 days, a cow must conceive within 80 days postpartum to maintain a yearly calving interval. In lactating heifers and cows beginning at the second or greater parity, milk output generally peaks at about 2 months postpartum, but feed intake usually lags behind, resulting in negative energy balance during early to midlactation (Bauman and Currie, 1980). The low plane of nutrition will be more evident if cows are managed to calf during the dry or winter seasons when poor forage quality is more probable. For example, the average weight loss by the University of Wyoming spring-calving cowherd was 11.8 kg while grazing native range during mid-summer. Only 54% of the cows were losing weight, and average weight lost by those cows in negative energy balance was 31.7 kg. The latter average weight loss is comparable to average weight loss experienced by a fall-calving cowherd grazing native range during the autumn (30.8 kg). With the aforementioned loss of body weight during early to mid-gestation in mind, researchers at the University of Wyoming Center for the Study of Fetal Programming conducted a study in which Angus x Gelbvieh rotationally crossed cows were blocked by body weight and allocated to either a control or nutrient restricted diet from day 31 to 120 of gestation. The control diet consisted of native grass hay (12.1% CP; 70.7% in vitro OM disappearance [IVOMD]) fortified with vitamins and minerals fed at NRC (1996) recommendations for a non-lactating, mature cow to gain 0.72 kg/day during the initial 120 day of gestation. The nutrient-restricted diet consisted of feeding one-half of the control diet‟s minerals and vitamins and millet straw (9.9% CP; 54.5% IVOMD) to provide 68.1% NE m and 86.7% of metabolizable protein requirements during the first 120 day of gestation (NRC, 1996). At day 125 of gestation, a subset of cows was harvested for tissue collection, and the remaining were blocked by treatment, body weight, and body condition score, and allocated to 16 pens. Control cows remained on the control diet, and nutrient restricted cows were provided a realimentation diet consisting of the nutrient restricted hay and the control diet‟s vitamins and minerals, plus a supplement of 79.6% cracked corn, 6.1% soybean meal, 5.3% sunflower meal, 4.2% cane molasses, 2.6% safflower seed meal, and 1.6% dried skim milk (13.2% CP; 77.6% IVOMD) to provide 2.15 Mcal more NEm/day than the control diet to help nutrient restricted cows attain a body condition score similar to the control cows by day 250 of gestation (Figure 3). Another subset of cows was harvested for tissue collection at day 250 of gestation, and the remaining cows were then co-mingled and fed to meet NRC (1996) requirements for late gestation beef cows. 111 from birth to weaning then backgrounded to the same feedlot entry weight. Thus, limiting compensatory growth by nutritional restriction up to weaning resulted in smaller cattle and carcasses at an equivalent age end points. Effects of early-postpartum maternal nutrition on growth, feed intake, and growth efficiency of steers (Stalker et al., 2006) and heifers (Martin et al., 2007) in the feedlot soon after weaning were not evident. Literature reviewed by Berge (1991) also demonstrated that feed conversion efficiency was not necessarily affected by plane of nutrition before weaning. As discussed earlier, an assessment of the relationship between fetal development and subsequent postnatal performance requires consideration of the consequences of nutrition during pregnancy on subsequent maternal performance when offspring rely on their dams for the majority of their nutrients. This concept can be illustrated by evaluation of an ongoing study being conducted at the University of Wyoming (Price et al., 2007). On day 45 of gestation, 36 of most uniform cows (12 triparous and 24 diparous) were selected to be individually fed native grass hay plus 1 of 3 supplements from day 45 through 185 of gestation. The control diet consisted of native grass hay plus a soybean meal-based supplement formulated for pregnant replacement heifers (590 kg mature BW) to achieve 0.43 kg/day of BW gain (NRC, 2000), which we estimated would be comparable to daily gain of 0.51 kg/day of BW gain for non-lactating cows pregnant with their second and third calf. The other dietary treatments were 70% of NEm provided by control, and 70% of NE m provided by control plus a ruminally undegraded protein supplement (6.8% porcine blood meal, 24.5% hydrolyzed feather meal, and 68.7% menhaden fish meal; DM basis; Scholljegerdes et al., 2005a) designed to provide duodenal essential amino acid flow equal to that of cattle fed control. Basal flows of total essential amino acids were predicted for each treatment using the equation reported by Scholljegerdes et al. (2004; total essential amino acid flow to the small intestine, g/d = [0.055 × g of OM intake] + 1.546). The quantity of protein supplement delivered was adjusted for more extensive ruminal degradation of the supplement in cattle fed restricted amounts of forage (Scholljegerdes et al., 2005b). Using actual mineral content of each dietary ingredient, the control supplement was fortified to ensure that cows fed control would consume the same amount of mineral per unit body weight as cows fed the protein supplement. Feed offered daily was adjusted for biweekly changes in body weight and increased NEm requirements as gestation proceeded. Control cows had greater body weight than nutrient restricted cows from day 73 through 185 of gestation. Body weight of cows fed the protein supplement was intermediate until day 115 of gestation; these cows had significantly greater body weight than nutrient restricted cows thereafter, but body weight of cows fed supplemental protein did not differ from control cows throughout the experiment. Change in body weight for nutrient restricted cows was not statistically significant indicating that the nutrient restricted cows maintained body weight. Cows were fed to achieve similar body weight within 4 weeks of parturition, and calf birth weights were not affected by gestational dietary subsequent carcass merits are reduced if steers were gestated by cows gaining 17 kg less during mid-gestation. Consequences of Poor Nutrition during the Latter Stages of Gestation Offspring born with low birth weights are less viable and adjust less rapidly to the extrauterine environment (Cundiff et al., 1986). Furthermore, low birth weight is associated with high neonatal morbidity and mortality rates. In reviewing 13 datasets in which researchers placed cows on a low plane of nutrition during late gestation, Dunn (1980) noted that the incidence of neonatal calf mortality increased 5 percentage units if cows were underfed. More recent data summarized by Wu et al. (2006) showed that preweaning deaths of heifer calves born alive in the United States was 10.5% with 70% of the deaths occurring within the first 7 days after birth. Incidence of neonatal mortalities may be linked to less thriftiness and reduced immune transfer for calves born to cows that were undernourished during gestation (Odde, 1988). If the low birth weight calf survives the early neonatal period, it is possible that it will grow slower than calves of normal birth weight at all stages of postnatal growth. Calves born 35% lighter due to severely restricted maternal nutrition from day 80 to 90 of pregnancy to parturition remained smaller at any given postnatal age compared with well-grown or better nourished counterparts (Greenwood and Cafe, 2007). However, it is difficult to ascertain whether or not this response represents permanent stunting or simply delayed of attainment of mature size. When differences in birth weight were less severe, however, postweaning growth was not significantly affected by birth weight (Cafe et al., 2006a). Nonetheless, influences of nutrition during mid and late pregnancy on calf weaning weight have been demonstrated, irrespective of effects on fetal growth (Cafe et al., 2006b; Stalker et al., 2006). Although effects of variable nutrition during mid and (or) late pregnancy on weight at birth seem to be overcome by adequate nutrition within 56 days postpartum (Freetly et al., 2000), a cursory review of several datasets would suggest that, while not statistically significant in each experiment, absolute weight of cattle at the time of harvest is consistently less if calves are born lighter or are lighter at weaning (Freetly et al., 2000; Banta et al., 2006; Stalker et al., 2006). Postnatal Nutrition The major nutritional factors affecting pre-weaning calf growth and composition at weaning are the lactational performance of the dam and the quality and availability of nutrients from pasture and (or) supplementation before and after parturition. In summarizing data of Greenwood et al. (2005; 2006), Greenwood and Cafe (2007) illustrated that a difference in weaning weight of 73 kg resulted in a 40 kg difference in live weight and 24 kg in carcass weight at 30 months of age. The low weaning weight cattle grew more rapidly during backgrounding and at a similar rate in the feedlot. Similarly, Cafe et al. (2006a) reported some compensatory growth occurred in steers restricted in growth 112 Armstrong, D. G., J. G. Gong, and R. Webb. 2003. Interactions between nutrition and ovarian activity in cattle: Physiological, cellular and molecular mechanisms. Reprod. Suppl. 61:403–414. Armstrong, D. G., T. G. McEvoy, G. Baxter, J. J. Robinson, C. O. Hogg, K. J. Woad, R. Webb, and K. D. Sinclair. 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. Baker, D. H. 2008. Animal models in nutrition research. J. Nutr. 138:391-396. Banta, J. P., D. L. Lalman, F. N. Owens, C. R. Krehbiel, and R. P. Wettemann. 2006. Effects of interval-feeding whole sunflower seeds during mid to late gestation on performance of beef cows and their progeny. J. Anim. Sci. 84:2410-2417. Baracos, V. E. 2004. Animal models of amino acid metabolism: a focus on the intestine. J. Nutr. 134:1656S-1659S. Barcroft, J. 1946. Researches on Pre-Natal Life. Blackwell Scientific Publications, Oxford. Barker, D. J. P. 1994. Mothers, Babies and Disease in Later Life. BMJ Publishing Group. London. Barker, D. J. P. 2004. Developmental origins of well being. Philosiol. Transl. Royal Soc. (London) 359:1359-1366. Barker, D. J. P., K. M. Godfrey, C. Osmond, and A. Bull, 1992. The relation of fetal length, ponderal index and headcircumference to blood pressure and the risk of hypertension in adult life. Paediatric Perinatal Epidemiol. 6:35–44. Barker, D. J. P., P. D. Gluckman, K. M. Godfrey, J. E. Harding, J. A. Owens, and J. S. Robinson. 1993. Fetal nutrition and cardiovascular disease in adult life. Lancet 2341:938-941. Barker, D. J., P. D. Winter, C. Osmond, B. Margetts, and S. J. Simmonds. 1989. Weight in infancy and death from ischaemic heard disease. Lancet 2:577-580. Barry, J. S. and R. V. Anthony. 2008. The pregnant sheep as a model for human pregnancy. Theriogenol. 69:5567. Bauman D. E., J. H. Eisemann, and W. B. Currie. 1982. Hormonal effects on partitioning of nutrients for tissue growth: role of growth hormone and prolactin. Fed. Proc. 41:2538-2544. Bauman, D. E. and W. B. Currie. 1980. Partitioning of nutrients during pregnancy and lactation: A review of mechanisms involving homeostasis and homeorhesis. J. Dairy Res. 63:1514-1529. Bauman, D. E., J. W. Perfield II, K. J. Harvtine, and L. H. Baumgard. 2008. Regulation of fat synthesis by conjugated linoleic acid: lactation and the ruminant model. J. Nutr. 138:403-409. Beaty, J. L., R. C. Cochran, B. A. Lintzenich, E. S. Vanzant, J. L. Morrill, R. T. Brandt, Jr., and D. E. Johnson. 1994. Effect of frequency of supplementation and protein concentration in supplements on performance and digestion characteristics of beef cattle consuming low-quality forages. J. Anim. Sci. 72:24752486. treatment. Despite common postpartum nutritional management, cows fed supplemental protein during early to mid-gestation produced 1.7 kg/d less milk at peak lactation than cows from the other two gestational dietary treatments. As a result, calves suckling cows fed supplemental protein during gestation gained 7.1 kg/d less than calves suckling cows from the other gestational dietary treatments. Although the long-term consequences of a lesser growth rate during the suckling phase are still under investigation, results of an extensive literature review by Berge (1991) would suggest that calves are not likely to compensate for their lower plane of nutrition imposed during the first 6 months of life. SUMMARY AND CONCLUSIONS Accumulating data is providing traction for relevance of developmental programming concepts in livestock. Maternal nutrition during gestation is a major determinant to fetal growth. Improper nutrition during gestation creates situations during critical windows of development where offspring may be predisposed to long-term complications. Inadequate nutrition during early to mid-gestation alters development of a variety of fetal tissues, which may affect subsequent health but effects on postnatal growth performance and carcass characteristics seem to be equivocal. A low plane of nutrition during mid to late gestation may reduce the ruminant offspring‟s absolute body weight at birth through slaughter, but again, effects on postnatal growth performance and carcass characteristics are less affected if gestation nutrition restriction is not severe. Improper nutrition during the perinatal period will increase offspring mortality. A low plane of nutrition during lactation may reduce absolute body weight throughout the growing period because calves experiencing lower planes of nutrition during the first 6 months of life are less apt to compensate later in life. Additional research efforts in these areas are needed to quantify degree of impact of maternal nutritional perturbations during gestation on both short and long-term metabolism, health, production, product quality, and economics in ruminant livestock species. LITERATURE CITED Anthony, R. V., R. A. Bellows, R. E. Short, R. B. Staigmiller, C. C. Kaltenbach, and T. G. Dunn. 1986. Fetal growth of beef calves. II. Effect of sire on prenatal development of the calf and related placental characteristics. J. Anim. Sci. 62:1375-1387. Armitage, J. A, I. Y. Khan, P. D. Taylor, P. W. Nathanielsz, and L. Poston. 2004. Developmental programming of the metabolic syndrome by maternal nutritional imbalance: how strong is the evidence from experimental models in mammals? J. Physiol. 561:355377 Armitage, J. A., P. D. Taylor, and L. Poston. 2005. Experimental models of developmental programming: consequences of exposure to an energy rich diet during development. J. Physiol. 565:3-8. 113 Bedi, K. S., A. R. Birzgalis, M. Mahon, J. L. Smart, and A. C. Wareham. 1982. Early life undernutrition in rats. 1. Quantitative histology of skeletal muscles from underfed young and refed adult animals. Br. J. Nutr. 47:417–431. Bee, G. 2004. Effect of early gestation feeding, birth weight, and gender of progeny on muscle fiber characteristics of pigs at slaughter. J. Anim. Sci. 82:826–836. Bellows, R. A., and R. E. Short. 1978. Effects of precalving feed level on birth weight, calving difficulty and subsequent fertility. J. Anim. Sci. 46:1522-1528. Bellows, R. A., R. E. Short, D. C. Anderson, B. W. Knapp, and O. F. Pahnish. 1971. Cause and effect relationships associated with calving difficulty and calf birth weight. J. Anim. Sci. 33:407-415. Berge, P. 1991. Long-term effects of feeding during calfhood on subsequent performance of beef cattle (a review). Livest. Prod. Sci. 28:179-201. Bergen, W. G. 2007. Contribution of research with farm animals to protein metabolism concepts: a historical perspective. J. Nutr. 137:706-710. Bergen, W. G. and H. J. Mersmann. 2005. Comparative aspects of lipid metabolism: impact on contemporary research and use of animal models. J. Nutr. 135:24992502. Berry, J. S. and R. V. Anthony. 2008. The pregnant sheep as a model for human pregnancy. Theriogenol. 69:5567. Bispham, J., G. S. Gopalakrishnan, J. Dandrea, V. Wilson, H. Budge, D. H. Keisler, F. Broughton Pipkin, T. Stephenson, and M. E. Symonds. 2003. Maternal endocrine adaptation throughout pregnancy to nutritional manipulation: consequences for maternal plasma leptin and cortisol and the programming of fetal adipose tissue development. Endocrinol. 144:35753585. Bloomfield, F. H., M. H. Oliver, P. Hawkins, M. Campbell, D. J. Phillips, P. D. Gluckman, J. R. G. Challis, and J. E. Harding. 2003. A periconceptional nutritional origin for noninfectious preterm birth. Science 330:606. Boland, M. P. and P. Lonergan. 2005. Effects of nutrition on fertility in dairy cows. Adv. Dairy Techn. 15:19–33. Boland, M. P. and P. Lonergan. 2005. Effects of nutrition on fertility in dairy cows. Adv. Dairy Techn. 15:19–33. Boland, M. P., P. Lonergan, and D. O‟Callaghan. 2001. Effect of nutrition on endocrine parameters, ovarian physiology, and oocyte and embryo development. Theriogenol. 55:1323–1340. Bonora, D., S. Kiechl, J. Willeit, F. Oberhollenzer, G. Egger, R. C. Bonadonna, and M. Muggeo. 2003. Metabolic syndrome: epidemiology and more extensive phenotyic description. Cross-sectional data from the Bruneck Study. Int. J. Obes. Relat. Metab. Disord. 27:1283-1289. Borowczyk, E. D. A. Redmer, J. S. Caton, J. J. Bilska, R. M. Weigl, D. Rouse, K. A. Vonnahme, P. P. Borowicz, L. P. Reynolds, and A. T. Grazul-Bilska. 2006. Effects of plane of nutrition on in vitro fertilization (IVF) and early embronic development in sheep. J. Anim. Sci. 84:1593-1599. Bull, Sleeter and W. E. Carroll. 1937. Principles of Feeding Farm Animals. The MacMillan Company, New York. Cafe, L. M., D. W. Hennessy, H. Hearnshaw, S. G. Morris, and P. L. Greenwood. 2006a. Influences of nutrition during pregnancy and lactation on birth weights and growth to weaning of calves sired by Piedmontese or Wagyu bulls. Aust. J. Exp. Agric. 46:245-255. Cafe, L. M., H. Hearnshaw, D. W. Hennessy, and P. L. Greenwood. 2006b. Growth and carcass characteristics at heavy market weights of Wagyu-sired steers following slow or rapid growth to weaning. Aust. J. Exp. Agric. 46:951-955. Carlson, D. B., J. J. Reed, P. P. Borowicz, J. B. Taylor, L. P. Reynolds, T. L. Neville, D. A. Redmer, K. A. Vonnahme, and J. S. Caton. 2009. Effects of dietary selenium supply and timing of nutrient restriction during gestation on maternal growth and body composition of pregnant adolescent ewes. J. Anim. Sci. 87:669-680. Carpenter, K. J. 2003a. A short history of nutritional science: Part 1 (1785-1885). J. Nutr. 133:638-645. Carpenter, K. J. 2003b. A short history of nutritional science: Part 2 (1885-1912). J. Nutr. 133:975-984. Carpenter, K. J. 2003c. A short history of nutritional science: Part 3 (1912-1944). J. Nutr. 133:3023-3032. Carpenter, K. J. 2003d. A short history of nutritional science: Part 4 (1945-1985). J. Nutr. 133:3331-3342 Caton, J. S., J. J. Reed, R. Aitken, J. S. Milne, P. P. Borowicz, L. P. Reynolds, D. A. Redmer, and J. M. Wallace. 2009. Effects of maternal nutrition and stage of gestation on body weight, visceral organ mass, and indices of jejunal cellularity, proliferation, and vascularity in pregnant ewe lambs. J. Anim. Sci. 87:222-235. Caton, J., K. Vonnahme, J. Reed, T. Neville, C Effertz, C. Hammer, J. Luther, D. Redmer, and L. Reynolds. 2007. Effects of maternal nutrition on birth weight and postnatal nutrient metabolism. Proc. Interntl. Symp. Energy and Protein Metab. Vichy, France. EAAP Publication No. 124:101-102. Clanton, D. C. and D. R. Zimmerman. 1970. Symposium on pasture methods for maximum production in beef cattle: Protein and energy requirements for female beef cattle. J. Anim. Sci. 30:122-132. Clarke, L., D. S. Buss, D. T. Juniper, M. A. Lomax, and M. E. Symonds. 1997. Adipose tissue development during early postnatal life in the ewe-reared lamb. Exp. Physiol. 82:1015-1027. Cline, H. B., B. W. Neville, G. P. Lardy, and J. S. Caton. 2009. Influence of advancing season on dietary composition, intake, site of digestion, and microbial efficiency in beef steers grazing a native range in western North Dakota. J. Anim. Sci. 87: 375-383. Close W. H. and J. F. Pettigrew. 1990. Mathematical models of sow reproduction. J. Reprod. Fertil. Suppl. 40:83-88. Crampton, E. W. and L. E. Lloyd. 1959. Fundamentals of Nutrition. W. H. Freeman and Company. San Francisco. 114 Glore S. R. and D. K.Layman. 1983. Cellular growth of skeletal muscle in weanling rats during dietary restrictions. Growth 47:403-410. Gluckman, P. D. and M. A. Hanson. 2004a. The developmental origins of the metabolic syndrome. Trends Endocrinol. Metab. 15:183-187. Gluckman, P. D. and M. A. Hanson. 2004b. Living with the past: evolution, development, and patterns of disease. Science 305:1733-1736. Gnanalingham, M. G., A. Mostyn, M. E. Symonds, and T. Stephenson. 2005. Ontogeny and nutritional programming of adiposity in sheep: potential role of glucocorticoid action and uncoupling protein-2. Amer. J. Physiol. Regul. Integr. Comp. Physiol. 289:R1407R1415. Godfrey, K. M. and D. J. Barker. 2000. Fetal nutrition and adult disease. Am. J. Clin. Nutr. 71(Suppl.):1344S1352S. Grazul-Bilska, A. T., E. Borowczyk, W. Arndt, J. Evoniuk, M. O‟neil, J. J. Bilski, R. M. Weigl, J. D. Kirsch, K. C. Kraft, K. A. Vonnahme, D. A. Redmer, L. P. Reynolds, and J. S. Caton. 2006. Effects of overnutrition and undernutritionn on invitro fertilization (IVF) and early embryonic development in sheep. Biol. Repro. Special Issue, p. 131, Abstr. No. 267, 39th Ann. Soc. Study Reprod. Meeting. Greenwood, P. L. and L. M. Cafe. 2007. Prenatal and preweaning growth and nutrition of cattle: long-term consequences for beef production. Anim. 1:1283-1296. Greenwood, P. L. L. M. Cafe, H. Hearnshaw, D. W. Hennessy, and S. G. Morris. 2006. Long-term consequences of birth weight and growth to weaning for carcass, yield and beef quality characteristics of Piedmontese- and Wagyu-sired cattle. Aust. J. Exp. Agric. 46:257-269. Greenwood, P. L., A. S. Hunt, J. W. Hermanson, and A. W. Bell. 1998. Effects of birth weight and postnatal nutrition on neonatal sheep: I. Body growth and composition, and some aspects of energetic efficiency. J. Anim. Sci. 76:2354-2367. Greenwood, P. L., A. S. Hunt, J. W. Hermanson, and A. W. Bell. 2000. Effects of birth weight and postnatal nutrition on neonatal sheep: II. Skeletal muscle growth and development. J. Anim. Sci. 78:50-61. Greenwood, P. L., L. M. Cafe, H. Hearnshaw, and D. W. Hennessy. 2005. Consequences of nutrition and growth retardation early in life for growth and composition of cattle and eating quality of beef. Recent Adv. Anim. Nutr. Aust. 15:183-195. Guan, H., E. Arany, J. P. van Beek, A. Chamson-Reig, S. Thyssen, D. J. Hill, and K. Yang. 2005. Adipose tissue gene expression profiling reveals distinct molecular pathways that define visceral adiposity in offspring of maternal protein-restricted rats. Amer. J. Physiol. Endocrinol. Metab. 288:E663-E673. Hales, C. N. and D. J. Barker. 1992. Type 2 (non-insulindependent) diabetes mellitus: the thrifty phenotype hypothesis. Diabetologia 35:595-601. Hales, C. N. and D. J. P. Barker. 2001. The thrifty phenotype hypothesis. Br. Med. Bull. 60:21–32. Cundiff, L. V., M. V. Macneil, and K. E. Gregory. 1986. Between- and within-breed analysis of calving and survival to weaning in beef cattle. J. Anim. Sci. 63:2733. Du, M. M. J. Zhu, W. J. Means, B. W. Hess, and S. P. Ford. 2004. Effect of nutrient restriction on calpain and calpastatin content of skeletal muscle from cows and fetuses. J. Anim. Sci. 82:2541-2547. Du, M. M. J. Zhu, W. J. Means, B. W. Hess, and S. P. Ford. 2005. Nutrient restriction differentially modulates the mammalian target of rapamycin signaling and the ubiquitin-proteasome system in skeletal muscle of cows and their fetuses. J. Anim. Sci. 83:114-123. Dunn, T. G. 1980. Relationship of nutrition to successful embryo transplantation. Theriogenol. 13:27-39. Fernandez, X. and E. Thornbert. 1991. A review of the causes of variation in muscle glycogen content and ultimate pH in pigs. J. Muscle Foods 2:209-235. Ferrell, C. L., S. P. Ford, R. L. Prior, and R. K. Christenson. 1983. Blood flow, steroid secretion and nutrient uptake of the gravid bovine uterus and fetus. J. Anim. Sci. 56:656-667. Ferrell, C. L., W. N. Garrett, and N. Hinman. 1976. Growth, development and composition of the udder and gravid uterus of beef heifers during pregnancy. J. Anim. Sci. 42:1477-1489. Ford, S. P., B. E. Burt, B. W. Hess, M. Du, P. W. Nathanielsz, and M. J. Nijland. 2005. Effects of early maternal undernutrition and realimentation on kidney development in the bovine fetus. Proc. Soc. Gynecol. Invest. Los Angeles, CA. Ford, S. P., B. W. Hess, M. M. Schwope, M. J. Nijland, J. S. Gilbert, K. A. Vonnahme, W. J. Means, H. Han, and P. W. Nathanielsz. 2007. Maternal undernutrition during early to mid- gestation in the ewe results in altered growth, adiposity, and glucose tolerance in male offspring. J. Anim. Sci. 85:1285-1294. Fowden, A. L., D. A. Giussani, and A. J. Forhead. 2006. Intrauterine programming of physiological systems: Causes and consequences. Physiol. 21:29-37. Freetly, H. C., C. L. Ferrell, and T. G. Jenkins. 2000. Timing of realimentation of mature cows that were feed-restricted during pregnancy influences calf birth weights and growth rates. J. Anim. Sci. 78:2790-2796. Gardner, B. A., H.G. Dolezal, F. N. Owens, L. L. K. Bryant, J. L. Nelson, B. R. Schutte, and R. A. Smith. 1998. Impact of health on profitability of feedlot steers. Pages 102-108 in Okla. State Univ. Anim. Sci. Res. Rep. Gardner, D. S., K. Tingey, B. W. M. Van Bon, S. E. Ozanne, V. Wilson, J. Dandrea, D. H. Keisler, T. Stephenson, and M. E. Symonds. 2005. Programming of glucose-insulin metabolism in adult sheep after maternal nutrition. Amer. J. Physiol. Regul. Integr. Comp. Physiol. 289:R947-R954. Gilbert, J. S., A. L. Lang, A. R. Grant, and M. J. Nijland. 2005. Maternal nutrient restriction in sheep: hypertension and decreased nephron number in offspring at 9 months of age. J. Physiol. 565.1:137-147. 115 Han, H., T. R. Hansen, B. Berg, B. W. Hess, and S. P. Ford. 2008. Maternal undernutrition induces differential cardiac gene expression in pulmonary hypertensive steers at high elevation. Am. J. Physiol. Heart Circ. Physiol. 295:H382-H389. Hanson, M. A, and P. D. Gluckman. 2005. Developmental processes and the induction and cardiovascular function: conceptual aspects. J. Physiol. 565:27–34. Hausman, D. B., M. DiGirolamo, T. J. Bartness, G. J. Hausman, and R. J. Martin. 2001. The biology of white adipocyte proliferation. Obes. Rev. 2:239-254. Hess, B. W., S. L. Lake, E. J. Scholljegerdes, T. R. Weston, V. Nayigihugu, J. D. C. Molle, and G. E. Moss. 2005. Nutritional controls of beef cow reproduction. J. Anim. Sci. 83:E90-E106. Holland, M. D. and K. G. Odde. 1992. Factors affecting calf birth weight: A review. Theriognol. 38:769-798. Johnson, D. E. 2007. Contributions of animal nutrition research to nutritional principles: energetics. J. Nutr. 137:698-701. Johnson, J. A., J. S. Caton, W. W. Poland, D. R. Kirby, and D. V. Dhuyvetter. 1998. Influence of season dietary composition, intake, and digestion by beef steers grazing mixed-grass prairie in western North Dakota. J. Anim. Sci. 76:1683-1690. Kendrick, K. W., T. L. Bailey, A. S. Garst, A. W. Pryor, A. Ahmadzadeh, R. M. Akers, W. E. Eyestone, R. E. Pearson, and F. C. Gwazdauskas. 1999. Effects of energy balance of hormones, ovarian activity, and recovered oocytes in lactating Holstein cows using transvaginal follicular aspiration. J. Dairy Sci. 82:1731– 1741. Kumarasamy, V., M. D. Mitched, F. H. Bloomfield, M. H Oliver, M. E. Cambell, J. R. Challis, and J. E. Harding. 2005. Effects of periconceptional undernutrition on the initiation of parturition in sheep. Amer. J. Physiol. Regul. Integr. Comp. Physiol. 288:R67-R72. Lake, S. L., E. J. Scholljegerdes, R. L. Atkinson, V. Nayigihugu, S. I. Paisley, D. C. Rule, G. E. Moss, T. J. Robinson, and B. W. Hess. 2005. Body condition score at parturition and postpartum supplemental fat effects on cow and calf performance. J. Anim. Sci. 83:29082917. Lardy, G. P. and J. S. Caton. 2010. Beef Cattle Nutrition in Commercial Ranching Systems. In: Ruminant Nutrition Production Systems. UNESCO Books. (Accepted). Larson, D. M., J. L. Martin, D. C. Adams, and R. N. Funston. 2009. Winter grazing system and supplementation during late gestation influence performance of beef cows and steer progeny. J. Anim. Sci. 87:1147-1155. Luther J. S., D. A. Redmer, L. P. Reynolds, and J. M. Wallace. 2005. Nutritional paradigms of ovine fetal growth restriction: implications for human pregnancy. Human Fertility. 8:179-187. Magness, R. R. 1998. Maternal cardiovascular and other physiological responses to the endocrinology of pregnancy. Pages 507-539 In: The Endocrinology of Pregnancy. F.W. Bazer (ed) Humana Press, Totowa, NJ. Maltin, C. A., C. C. Warkup, K. R. Matthews, C. M. Grant, A. D. Porter, and M. I. Delday. 1997. Pig muscle fibre characteristics as a source of variation in meat quality. Meat Sci. 47:237-248. Martin, J. L., K. A. Vonnahme, D. C. Adams, G. P. Lardy, and R. N. Funston. 2007. Effects of dam nutrition on growth and reproductive performance of heifers calves. J. Anim. Sci. 85:841-847. Martin-Gronert, M. S. and S. E. Ozanne. 2006. Maternal nutrition during pregnancy and health of the offspring. Biochem. Soc. Trans. 34:779-782. Mathews, D. M. 1991. Protein absorption: development and present state of the subject. John Wile & Sons, Inc. New York, NY. Maynard, L. A. and J. K. Loosli. 1956. Animal Nutrition. McGraw-Hill Book Company, Inc. McCollum, E. V. 1957. A History of Nutrition. The Riversid Press. Cambridge MA. McDowell, L. R. 2003. Minerals in Animal and Human Nutrition 2nd ed. Elsevier Press. Amsterdam. McDowell. L. R. 2000. Vitamins in animal and human nutrition. Iowa State Univ. Press. Ames. McEvoy, T. G., J. J. Robinson, R. P. Aitken, P. A. Findlay, R. M. Palmer, and I. S. Robertson. 1995. Dietaryinduced suppression of pre-ovulatory progesterone concentrations in superovulated ewes impairs the subsequent in vivo and in vitro development of their ova. Anim. Reprod. Sci. 39:89–107. Mellor D. J. 1983. Nutritional and placental determinants of foetal growth rate in sheep and consequences for the newborn lamb. Br. Vet. J. 139:307-324. Mersmann, H. J. and S. B. Smith. 2005. Development of white adipose tissue lipid metabolism. Pages 275-302 In: Biology of Metabolism in Growing Animals. D. G. Burrin and H. J. Mersmann (ed) Elsevier, Philadelphia, PA. Meschia G. 1983. Circulation to female reproductive organs. Pages 241-267 In: Handbook of Physiology Vol. 3. Amer. Physiol. Soc. Bethesda, MD. Meyer, A. M., J. J. Reed, K. A. Vonnahme, S. A. SotoNavarro, L. P. Reynolds, S. P. Ford, B. W. Hess, and J. S. Caton. 2010a. Effects of stage of gestation and nutrient restriction during early to mid-gestation on maternal and fetal visceral organ mass and indices of jejunal growth and vascularity in beef cows. J. Anim. Sci. doi:10.2527/jas.2009-2220. Meyer, A. M., J. J. Reed, T. L. Neville, J. B. Taylor, C. J. Hammer, L. P. Reynolds, D. A. Redmer, K. A. Vonnahme, and J. S. Caton. 2010b. Effects of nutritional plane and selenium supply during gestation on ewe and neonatal offspring performance, body composition, and serum selenium. J. Anim. Sci. 88:1786-1800. Miller, W. H., Jr., I. M. Faust, and J. Hirsch. 1984. Demonstration of de novo production of adipocytes in adult rats by biochemical and radioautographic techniques. J. Lipid Res. 25:336-347. Mitchel, A. D. 2007. Impact of research with cattle pigs and sheep on nutritional concepts: body composition and growth. J. Nutr. 137:711-714. 116 O‟Callaghan, D., H. Yaakub, P. Hyttel, L. J. Spicer, and M. P. Boland. 2000. Effect of nutrition and superovulation on oocyte morphology, follicular fluid composition and systemic hormone concentrations in ewes. J. Reprod. Fertil. 18:303–313. O‟Dell, B. L. and R. A. Sunde. 1997. Handbook of nutritionally essential mineral elements. Marcel Dekker, Inc. New York, NY. Odde, K. G. 1988. Survival of the neonatal calf. Vet. Clin. N. Amer. Food Anim. Pract. 4:50-58. Ouali, A. 1990. Meat tenderization: possible causes and mechanisms. A review. J. Muscle Foods. 1: 129-165. Papadopoulos, S., P. Lonergan, V. Gath, K. M. Quinn, A. C. Evans, D. O‟Callaghan, and M. P. Boland. 2001. Effect of diet quantity and urea supplementation on oocyte and embryo quality in sheep. Theriogenol. 55:1059–1069. Parr, R. A., A. H. Williams, I. P. Campbell, G. F. Witcome, and A. M. Roberts. 1986. Low nutrition of ewes in early pregnancy and the residual effect on the offspring. J. Agric. Sci. (Camb.) 106:81-87. Perry, R. C., L. R. Corah, R. C. Cochran, W. E. Beal, J. S. Stevenson, J. E. Minton, D. D. Simms, and J. R. Brethour. 1991. Influence of dietary energy on follicular development, serum gonadotropins, and first postpartum ovulation in suckled beef cows. J. Anim. Sci. 69:37623773. Perry, V. E. A., S. T. Norman, J. A. Owen, R. C. W. Daniel, and N. Phillips. 1999. Low dietary protein during early pregnancy alters bovine placental development. Anim. Reprod. Sci. 55:13-21. Perry, V. E. A., S. T. Norman, R. C. W. Daniel, P. C. Owens, P. Grant, and V. J. Doogan. 2002. Insulin-like growth factor levels during pregnancy in the cow are affected by protein supplementation in the diet. Anim. Reprod. Sci. 72:1-10. Pinkney, J. H., C. D. A. Stehouver, S. W. Coppack, J. S. Yudkin. 1997. Endothelial dysfunction: cause of the Insulin Resistance Syndrome. Diabetes 46(Suppl 2):S9– S13. Price, P. L., M. Du, S. I. Paisley, V. Nayigihugu, J. D. Hess, and B. W. Hess. 2007. Beef cow performance in response to early through mid-gestational nutrient restriction and provision of supplementary ruminally undegradable protein. Proc. West. Sect. Amer. Soc. Anim. Sci. 58:134-138. Prior, R. L. and D. B. Laster. 1979. Development of the bovine fetus. J. Anim. Sci. 48:1546-1553. Ramsey, E.M. 1982. The Placenta, Human and Animal. New York: Praeger. Rasby, R. J., R. P. Wettemann, R. D. Geisert, L. E. Rice, and C. R. Wallace. 1990. Nutrition, body condition and reproduction in beef cows: Fetal and placental development, and estrogens and progesterone in plasma. J. Anim. Sci. 68:4267-4276. Redmer, D. A., J. M. Wallace, and L. P. Reynolds. 2004. Effect of nutrient intake during pregnancy on fetal and placental growth and vascular development. Domest. Anim. Endocrinol. 27:199-217. Molle, J. D. C., E. J. Scholljegerdes, S. L. Lake, V. Nayigihugu, R. L. Atkinson, L. R. Miller, S. P. Ford, W. J. Means, J. S. Caton, and B. W. Hess. 2004. Effects of maternal nutrient restriction during early to midgestation on cow and fetal visceral organ measurements. Proc. West. Sect. Amer. Soc. Anim. Sci. 55:405-409. Moloney, A. P., P. Allen, and W. J. Enright. 2002. Body composition and adipose tissue accretion in lambs passively immunized against adipose tissue. Livestock Prod. Sci. 74:165-174. Monin, G. and A. Ouali. 1992. Muscle differentiation and meat quality. Pages 89-157 In: Developments in Meat Science. R. Lawrie (ed) Elsevier Applied Science, London.. Morrison, Frank. 1949. Feeds and Feeding: A Handbook for the Student and Stockman 21st Ed. The Morrison Publishing Company. Mossman, H. W. 1987. Vertebrate Fetal Membranes. Rutgers Univ. Press, Brunswick, NJ. Muhlhausler, B. S., C. T. Roberts, J. R. McFarlane, K. G. Kauter, and I. C. McMillen. 2002. Fetal leptin is a signal of fat mass independent of maternal nutrition in ewes fed at or above maintenance energy requirements. Biol. Reprod. 67:493-499. Nafikov, R. A. and D. C. Beitz. 2007. Carbohydrate and lipid metabolism in farm animals. J. Nutr. 137:702-705. Nathanielsz, P. W. 2006. Animal models that elucidate basic principles of the developmental origins of adult diseases. ILAR J. 2006. 47:73-82. Neville T. L., J.S. Caton, C.J. Hammer, J. J. Reed, J.S. Luther, J. B. Taylor, D.A. Redmer, L.P. Reynolds, and K.A. Vonnahme. 2010. Ovine offspring growth and diet digestibility are influenced by maternal Se supplementation and nutritional intake level during pregnancy despite a common postnatal diet. J. Anim. Sci. 88:(Submitted). Nichols, B. L. and P. J. Reeds. 1991. History of nutrition: history and current status of research in human energy metabolism. J. Nutr. 121:1889-1890. Nissen, P.M., V.O. Danielson, P.F. Jorgensen, and N. Oksbjerg. 2003. Increased maternal nutrition of sows has no beneficial effects on muscle fiber number or postnatal growth and has no impact on the meat quality of the offspring. J. Anim. Sci. 81:3018-3027. Nolan, R., D. O‟Callaghan, R. T. Duby, P. Lonergan, and M. P. Boland. 1998. The influence of short-term nutrient changes on follicle growth and embryo production following superovulation in beef heifers. Theriogenol. 50:1263–1274. NRC 1985. Nutrient Requirements of Sheep. Natl. Acad. Sci. Washington D.C., U.S.A. NRC. 1970. Maternal Nutrition and the Course of Pregnancy. Natl. Acad. Sci. Washington, DC. NRC. 1996. Nutrient Requirements of Beef Cattle. 7 th Ed. Natl. Acad. Sci. Washington, DC. NRC. 2000. Nutrient Requirements of Beef Cattle. 7 th Rev. Ed. Natl. Acad. Sci. Washington, DC. 117 Reynolds, L. P., P. P. Borowicz, K. A. Vonnahme, M. L. Johnson, A. T. Grazul-Bilska, J. M. Wallace, J. S. Caton, and D. A. Redmer. 2005b. Animal models of placental angiogenesis. Placenta. 26:689-708. Robinson J. J, K. D.Sinclair, and T. G. McEvoy. 1999. Nutritional effects on foetal growth. Anim. Sci. 68:315331. Robinson, J. J. 1983. Nutrition of the Pregnant Ewe. Pages 111-131 In: Sheep Production. W. Haresign (ed) Butterworths, London. Robinson, J. J. 1990. Nutrition in the reproduction of farm animals. Nutr. Res. Rev. 3:961–966. Sandstead, H. H. and L. M. Klevay. 2000. History of nutrition symposium: trace element nutrition and human health. J. Nutr. 130:483S-484S. Schantz, P., E. Fox, and E. Hutchinson. 1983. Muscle fibre type distribution, muscle cross-sectional area and maximal voluntary strength in humans. Acta Physiologica Scandinavica 117:219-226. Scholljegerdes, E. J. T. R. Weston, P. A. Ludden, and B. W. Hess. 2005a. Supplementing a ruminally undegradable protein supplement to maintain essential amino acid supply to the small intestine when forage intake is restricted in beef cattle. J. Anim. Sci. 83:2151–2161. Scholljegerdes, E. J., P. A. Ludden, and B. W. Hess. 2004. Site and extent of digestion and amino acid flow to the small intestine in beef cattle consuming limited amounts of forage. J. Anim. Sci. 82:1146–1156. Scholljegerdes, E. J., P. A. Ludden, and B. W. Hess. 2005b. Effect of restricted forage intake on ruminal disappearance of bromegrass hay and a blood meal, feather meal and fish meal supplement. J. Anim. Sci. 83:2146–2150. Shillabeer, G. and D. C. Lau. 1994. Regulation of new fat cell formation in rats: the role of dietary fats. J. Lipid Res. 35:592-600. Sletmoen-Olson, K. E., J. S. Caton, L. P. Reynolds, and K. C. Olson. 2000. Undegraded intake protein supplementation: I. Effects on forage utilization and performance of periparturient beef cows fed low-quality hay during gestation and lactation. J. Anim Sci. 78:449455. Smith, G. C., J. W. Savell, H. G. Dolezal, T. G. Field, D. R Gill, D. B. Griffin, D. S. Hale, J. B. Morgan, S. L. Northcutt, and J. D. Tatum. 1995. The final report of the National Beef Quality Audit. Colorado State Univ. Fort Collins; Texas A&M Univ., College Station; and Oklahoma State Univ., Stillwater, OK. NCA, Eaglewood, CO. Spitzer, J. C., D. G. Morrison, R. P. Wettemann, and L. C. Faulkner. 1995. Reproductive responses and calf birth and weaning weight as affected by body condition at parturition and postpartum wieght gain in primiprous beef cows. J. Anim. Sci. 73:1251-1257. Spurlock, M. E. and N. K. Gabler. 2008. The developmental of porcine models of obesity and the metabolic syndrome. J. Nutr. 138:397-402. Stahl, C. H., X. Lei, and B. Larson. 2008. Introduction to the Symposium: appropriate animal models for nutritional research in health and disease. J. Nutr. 138:389-390. Reed, J. J., A. L. Gevlin, G. P. Lardy, M. L. Bauer, and J. S. Caton. 2006. Effect of creep feed supplementation and season on intake, microbial protein synthesis and efficiency: Ruminal fermentation, digestion, rate of passage and performance in nursing calves grazing native range in southeastern North Dakota. J. Anim Sci. 84:411-423. Reed, J. J., M. A. Ward, K. A. Vonnahme, T. L. Neville, S. L. Julius, P. P. Borowicz, J. B. Taylor, D. A. Redmer, A. T. Grazul-Bilska, L. P. Reynolds, and J. S. Caton. 2007. Effects of Selenium supply and dietary restriction on maternal and fetal body weight, visceral organ mass, cellularity estimates, and jejunal vascularity in pregnant ewe lambs. J. Anim. Sci. 85:2721-2733. Rehfeldt, C. and G. Kuhn. 2006. Consequences of birth weight for postnatal growth performance and carcass quality in pigs as related to myogenesis. J. Anim. Sci. 84: E113-E123. Rehfeldt, C., I. Fiedler, and N. C. Stickland. 2004. Number and size of muscle fibers in relation to meat production. Page 1 In: Muscle Development of Livestock Animals: Physiology, Genetics, and Meat Quality. M. F. W. te Pas, M. E. Haagsman, and H. P. Everts (ed) CAB Int., Wallingford, UK. Reynolds L. P., P. P Borowicz, J. S. Caton, K. A. Vonnahme, J. S. Luther, D. S. Buchanan, S. A. Hafez, A. T. Grazul-Bilska, and D. A. Redmer. 2010a. Uteroplacental vascular development and placental function: An update. Invited review. Interntl. J. Develop. Biol. Special Issue „Placental Developmental Biology,” JS Hunt and KL Thornburg (eds) doi: 10.1387/ijdb.082799lr. Reynolds LP. 1986. Utero-ovarian interactions during early pregnancy: Role of conceptus-induced vasodilation. J Anim Sci. 62(Suppl. 2):47-61. Reynolds, L. P. and D. A. Redmer. 1995. Utero-placental vascular development and placental function. J. Anim. Sci. 73:1839-1851. Reynolds, L. P., D. S. Millaway, J. D. Kirsch, J. E. Infeld, and D. A. Redmer. 1990. Growth and in-vitro metabolism of placental tissues of cows from day 100 to day 250 of gestation. J. Reprod. Fert. 89:213-222. Reynolds, L. P., J. J. Ireland, J. S. Caton, D. E. Bauman, and Teresa A. Davis. 2009. Commentary on domestic animals in agricultural and biomedical research: An endangered enterprise. J. Nut. 139:427-28. Reynolds, L. P., J. S. Caton, D. A. Redmer, A. T. GrazulBilska, K. A. Vonnahme, P. P. Borowicz, J. S. Luther, J. M. Wallace, G. Wu, and T. E. Spencer. 2006. Evidence for altered placental blood flow and vascularity in compromised pregnancies. J. Physiol. 572:51-58. Reynolds, L. P., P. P. Borowicz, D. A. Vonnahme, M. L. Johnson, A.T. Grazul-Bilska, D. A. Redmer, and J. S. Caton. 2005a. Placental angiogenesis in sheep models of compromised pregnancy. J. Physiol. 565:43-58. Reynolds, L. P., P. P. Borowicz, J. S. Caton, K. A. Vonnahme, J. S. Luther, C. J. Hammer, K. R. Maddock Carlin, A. T. Grazul-Bilska, and D. A. Redmer. 2010b. Large animal models of developmental programming. J. Anim. Sci. 88:E61-E72. 118 Vonnahme, K.A., M.J. Zhu, P.P. Borowicz, T.W. Geary, B.W. Hess, L.P. Reynolds, J.S. Caton, W.J. Means, and S.P. Ford. 2007. Effect of early gestational undernutrition on angiogenic factor expression and vascularity in the bovine placentome. J. Anim. Sci. 85:2464-2472. Wallace J. M., D. A. Bourke, P. Da Silva, R. P. Aitken. 2001. Nutrient partitioning during adolescent pregnancy. Reprod. 122:347-357. Wallace L. R. 1948. The growth of lambs before and after birth in relation to the level of nutrition. J. Agric. Sci. (Camb.). 38:243-300 and 367-398. Wallace, J. M, R. P. Aitken, and M. A. Cheyne. 1996. Nutrient partitioning and fetal growth in rapidly growing adolescent ewes. J. Reprod. Fertil. 107:183190. Wallace, J. M., D. A. Bourke, and R. P. Aitken, 1999. Nutrition and fetal growth: paradoxical effects in the overnourished adolescent sheep. J. Reprod. Fertil. 54(Suppl. 1):385-399. Wallace, J. M., J. S. Milne, and R. P. Aitken. 2005. The effect of overnourishing singleton-bearing adult ewes on nutrient partitioning to the gravid uterus. Br. J. Nutr. 94:533-539. Wallace, L. R. 1948. The growth of lambs before and after birth in relation to the level of nutrition. J. Agric. Sci. (Camb.) 38:243-300. Ward, S. S. and N. C. Stickland. 1991. Why are slow and fast muscles differentially affected during prenatal undernutrition. Muscle Nerve 14:259–267. Webb, R., P. C. Garnsworthy, J.-G. Gong, and D. G. Armstrong. 2004. Control of follicular growth: Local interactions and nutritional influences. J. Anim. Sci. 82:E63–E74. Webb, R., R. G. Gosden, E. E. Telfer, and R. M. Moor. 1999. Factors affecting folliculogenesis in ruminants. Anim. Sci. 68:257–284. Wilson, P. W, and S. M. Grundy. 2003b. The metabolic syndrome: practical guide to origins and treatment: Part II. Circul. 108:1557-1540. Wilson, P. W. and S. M. Grundy. 2003a. The metabolic syndrome: practical guide to origins and treatment: Part I. Circul.108:1422-1424. Wu, G., F. W. Bazer, J. M. Wallace, and T. E. Spencer. 2006. Intrauterine growth retardation: implications for the animal sciences. J. Anim. Sci. 84:2316-2337. Yaakub, H., D. O‟Callaghan, and M. P. Boland. 1999. Effect of type and quantity of concentrates on superovulation and embryo yield in beef heifers. Theriogenol. 51:1259–1266. Zhu, M. J., M. Du, B. W. Hess, W. J. Means, P. W. Nathanielsz, and S. P. Ford. 2007. Maternal nutrient restriction upregulates signaling pathways in the cotyledonary artery of cow placentomes. Placenta 28:361-368. Zhu, M. J., M. Du, C. Tong, B. W. Hess, S. P. Ford, and P. W. Nathanielsz. 2009. Maternal overnutrition and obesity induced inflammatory response in fetal ileum of sheep. Biol. Reprod. 81:498. Stalker, A. L., D. C. Adams, T. J. Klopfenstein, D. M. Feuz, and R. N. Funston. 2006. Effects of pre- and postweaning nutrition on reproduction in spring calving cows and calf feedlot performance. J. Anim. Sci. 84:2582-2589. Stegeman, J. H. J. 1974. Placental development in the sheep and its relation to fetal development. Bijdragen Tot de dierkunde. 44:1-72. Stickland, N. C., S. Bayol, C. Ashton, and C. Rehfeldt. 2004. Manipulation of muscle fiber number. Page 69 In: Muscle Development of Livestock Animals: Physiology, Genetics, and Meat Quality. M. F. W. te Pas, M. E. Haagsman, and H. P. Everts (ed) CAB Int., Wallingford, UK. Swanson, T. J., C. J. Hammer, J. S. Luther, D. B. Carlson, J. B. Taylor, D. A. Redmer, T. L. Neville, J. J. Reed, L. P. Reynolds, J. S. Caton, and K. A. Vonnahme. 2008. Effects of gestational plane of nutrition and selenium supplementation on mammary development and colostrum quality in pregnant ewe lambs. J. Anim. Sci. 86:2415-2423. Symonds, M. E., S. Pearce, J. Bispham, D. S. Gardner, and T. Stephenson. 2004. Timing of nutrient restriction and programming of fetal adipose tissue development. Proc. Nutr. Soc. 63:397-403. Symonds, M. E., T. Stephenson, D. S. Gardner, H. Budge. 2007. Long-term effects of nutritional programming of the embryo and fetus: mechanisms and critical windows. Reprod. Fertil. Dev. 19:53-63. Taylor, P. D. and L. Poston. 2007. Developmental programming of obesity in mammals. Exp. Physiol. 92:287-298. Thomas, V. M. and R. W. Kott. 1995. A review of Montana winter range ewe nutrition research. Sheep and Goat Res. J. 11:17-24. Tripp, M. W., J. C. Ju, T. A. Hoagland, J. W. Riesen, X. Yang, and S. A. Zinn. 2000. Influence of somatotropin and nutrition on bovine oocyte retrieval and in vitro development. Theriogenol. 53:1581–1590. Underwood, E. J. and N. F. Suttle. 1999. The mineral nutrition of livestock. CABI Publishing. New York, NY. Underwood, K. R., B. W. Hess, C. M. Murrieta, D. C. Rule, S. I. Paisley, S. A. Thompson, S. P. Ford, and W. J. Means. 2008. Gestational maternal nutrient restriction effects on growth performance and carcass characteristics of steers. J. Anim. Sci. (Accepted with revisions). Vincent, I. C., H. L. I. Williams, and R. Hill. 1985. The influence of a low-nutrient intake after mating on gestation and perinatal survival of lambs. Br. Vet. J. 141:611-617. Vonnahme, K. A., B. W. Hess, T. R. Hansen, R. J. McCormick, D. C. Rule, G. E. Moss, W. J. Murdoch, M. J. Nijland, D. C. Skinner, P. W. Nathanielsz, and S. P. Ford. 2003. Maternal undernutrition from early to mid-gestation leads to growth retardation, cardiac ventricular hypertrophy and increased liver weight in the fetal sheep. Biol. Reprod. 69:133-140. 119 Zhu, M. J., S. P. Ford, P. W. Nathanielsz, and M. Du. 2004. Effect of maternal nutrient restriction in sheep on the development of fetal skeletal muscle. Biol. Reprod. 71:1968-1973. Table 1. Changes in fetal and placental weights, uterine and umbilical blood flows, and placental vascularity in various models of compromised pregnancy in sheep 1 Day of gestation2 Model Overfed adolescent High dietary Se 130 130-144 135 130 133-135 140 ↓2028% ↓17% ↓12% ↓11% ↓43% ↓42% ↓30% 135 NSE 130-134 Underfed adolescent Underfed adult Adolescent vs. Adult Genotype Heat-stressed adult Multiple pregnancy Fetal weight Placental weight Uterine blood flow Umbilical blood flow Vascularity3 ↓45% ↓36% ↓37% ↓31% (total capillary vol.) NSE --↓29% ↓47% ↓51% ↓37% ---↓17-32% ------↓26% ↓23% ---NSE ------↓60% ---- ↓24% ---- ---- ↓20% (cap. area density, CAR) ↓14% (cap. area density, CAR) ---↑36% ---↓30% (total cap. vol., COT) ↑20% (cap. number density, COT) ↑(cap. area density, CAR & COT) 4 Hypoxic (hypobaric) 140 NSE ---↓35% ---stress 1 Table adapted from Reynolds et al. (2006). 2 Length of gestation = approximately 145 d. 3 cap. = capillary; CAR = caruncle (maternal placenta); COT = cotyledon (fetal placenta/villus). 4 NSE = no significant effect. Table 2. Influence of maternal dietary selenium and nutritional intake during pregnancy on offspring body weight from birth through weaning at 57 d of age1 Selenium treatment2 ASe HSe Nutrition treatment3 CON HIGH P-values4 Nut Se*Nut Item SEM RES SEM Se BW, kg Birth Males 4.49 4.06 0.25 3.95a 4.78b 4.10a 0.29 0.18 0.06 0.96 a b Females 4.11 4.43 0.12 3.85 4.50 4.45b 0.15 0.05 0.002 0.71 Combined 4.25 4.37 0.10 4.02a 4.67b 4.25a 0.12 0.37 0.0007 0.66 Weaning, 57 d Males 19.52 19.45 0.95 18.10 20.35 20.01 1.27 0.96 0.28 0.75 Females 19.60 19.72 0.64 18.85 20.18 19.96 0.82 0.89 0.32 0.25 Combined 19.61 19.55 0.49 18.57a 20.33b 19.84ab 0.67 0.94 0.06 0.40 180 d Males 53.04 52.61 1.43 52.22ab 55.75a 50.51b 1.91 0.83 0.08 0.98 Females 49.41 51.47 1.73 50.59 50.80 49.94 2.15 0.35 0.95 0.85 Combined 51.23 51.98 1.11 51.31 53.27 50.24 1.48 0.61 0.26 0.80 1 Adapted from Neville et al. (2010). 2 Selenium treatments were daily intake of organically bound Se, adequate Se (ASe; 9.5 µg/kg BW) vs. high Se (HSe; 81.8 µg/kg BW). 3 Nutritional treatments were RES (fed at 60% of CON), CON (control; 100% requirements for gestating ewe lambs), and HIGH (fed at 140% of CON). 4 Probability values for effects of selenium (Se), nutrition (Nut), and the interaction. ab Means within a row having differing superscripts differ (P < 0.10). 120 Figure 1. Timeline of bovine fetal growth and development (courtesy of Dr. Kim Vonnahme). Periconceptional Implantation and Placental Growth Organogenesis Rapid Fetal Growth Perinatal Postnatal Conception Mid-Gestation Birth Figure 2. Critical windows of developmental programming (adapted from Fowden et al., 2006). 121 660 Weight (kg) 640 620 600 580 560 540 520 500 30 60 90 120 150 180 210 240 Day of Gestation Figure 3. Observed change in cow body weights in response to plane of nutrition from day 31 through 240 of gestation. Glomeruli per gram tissue 3500000 50000 3000000 * 2500000 * 2000000 40000 1500000 30000 1000000 500000 0 Glomeruli/gram tissue Absolute glomerular number Total glomerular numbers 20000 C NR Figure 4. Glomerular counts of kidneys collected from 250-day fetus of control (C) or nutrient restricted cows (NR). 122 MATERNAL PLANE OF NUTRITION: IMPACTS ON FETAL OUTCOMES AND POSTNATAL OFFSPRING RESPONSES J. S. Caton and B. W. Hess Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 POTENTIAL FOR NUTRITIONAL IMBALANCE IN HIGH-QUALITY FORAGES M. Kerley1 Division of Animal Sciences, University of Missouri, Columbia 65211 adoption by the industry. Change in metabolizable energy (ME) density of forages as plant growth proceeds has been well characterized. Management procedures (grazing management, legume interseeding, annuals, etc) have been developed to extend availability of high-quality forage to grazing cattle. Substantial periods of the growing season produce forage with ME density capable of supporting gains of growing calves in excess of 1 kg/d. Post-ruminal supply of AA (microbial and forage origins), however, are not sufficient to support energy-allowable gain and create an imbalance between energy and protein. We have concluded from our research that arginine, histidine, lysine, methionine, and threonine are potentially limiting, with the most limiting being methionine, in high-quality forage diets. Equally important is adequate degradable protein in forages containing less than 6% crude protein. This paper will focus on impact of an imbalanced absorbable AA to energy ratio on cow and weaned calf performance when fed a forage diet. INTRODUCTION Increased price of feed grains has provided a renaissance of interest in using forage for beef production post-weaning. Interest in maximizing use of forage for the cow herd has intensified as well due to economic pressures placed on beef producers. We find ourselves asking the question all businesses are asking, how we can do more with less? There are two topics emphasized in this paper: (1) how will selection for efficiency change cattle intake and (2) how nutrient imbalances can limit performance of forage-fed cattle. General agreement exists that forage intake is driven by nutrient demand. While several nutrient imbalances can influence intake, the two most readily at our control and likely to be limiting are energy and protein. A substantial body of research supports the hypothesis that intake is regulated by energy and AA requirement and supply. Rodent models have shown that satiety was controlled either when caloric or limiting AA consumption satisfied growth requirement. Extrapolating to cattle grazing forage or fed hay, it was hypothesized that intake and (or) performance would be determined by capability of the forage to satisfy nutrient requirements of the animal. Repeated observations of 1.4-fold difference in intake among populations of cattle with no difference in rate of gain led to the hypothesis that intake differences relate to differences in metabolic efficiency among individuals. This hypothesis was supported by research demonstrating equal correlative differences in mitochondrial respiration rates among individuals and also by the degree of heritability for efficiency. The importance of these findings is that feed efficiency can be theoretically improved 40% within beef populations via selection. Residual feed intake (RFI) populations are identified as efficient (1 standard deviation negative from the mean), inefficient (1 standard deviation positive from the mean), or average. Forage intake by nonlactating and lactating cows were reduced by 20 and 11%, respectively, when efficient cows were compared with inefficient cows (Meyer et al., 2008). As genetic selection skew cows in a herd toward greater efficiency, available forage should increase. This in turn should increase potential productivity of the animal, simultaneously creating greater potential for correcting nutritional imbalances. It is my view that economic value of feed efficiency technology will be pull-through for IMPACT OF ANIMAL EFFICIENCY SELECTION ON INTAKE Residual feed intake was introduced as a measure of efficiency to prevent confounding effects of growth rate that occurs with feed to gain ratio measurements. Residual feed intake is moderately heritable and has been biologically linked to mitochondrial metabolism. It is calculated as the difference between intake and expected intake. Expected intake is calculated by regressing intake of a population against their ADG and metabolic mid-weight (it is preferred to also include backfat into the regression). Independent variable regression coefficients are then used to calculate expected intake. Our interest was to measure the influence of selecting for efficiency (negative RFI) in a cow herd on forage intake. Fall calving Hereford cows were RFI phenotyped and then grouped as –RFI, average RFI, or +RFI. We contrasted –RFI against +RFI groups. Each group was assigned to 1 of 2 replicates based on BW and BCS. Cows grazed tall fescue (not infected with endophyte) pastures May to August, 2006 and February to April, 2007 (as pairs in 2007). Summer 2006 paddocks were grazed continuously. Forage availability was maintained in excess by using weekly electronic rising plate meter (RPM; FarmWorks, Feilding, New Zealand) readings to compute total RPM units per paddock. Rising plate meter units served as estimates of forage DM on offer, as each unit represented approximately the same DM yield on a given sampling date. Total RPM units per paddock were Key words: metabolic efficiency, metabolizable energy, rumen undegraded protein, intake 1 Corresponding author: [email protected] 125 computed to compare forage yield between paddocks in each replicate. This measure was corrected for paddock size using the equation: RPMtotal = RPMavg * Areapaddock. Paddocks were then adjusted in size to keep total RPM units within 10 per paddock for replicates from each pasture (block). Cows used during winter 2007 grazed on stockpiled tall fescue. Paddocks were strip-grazed using electric poly-tape and movable step-end posts, and a new strip was allocated every 3.5 d. Initial strip allocations were determined using a set residual and assuming that cows would consume 1.2% of their BW in NDF per day (Mertens, 1987). After this, strip size was allocated based upon the residual forage left. The goal of strip allocation was to not limit intake, while also keeping utilization similar between paddocks. Strip size was calculated based on the number of days expected for grazing each allocation when a 3.5 d moving schedule could not be kept. New grazing strips were provided as needed so that forage availability was not limiting for more than 12 h. Forage quality and availability were limiting at trial initiation for the winter grazing period; therefore, in order to maintain body condition of cows, cow-calf pairs were supplemented daily with 3.31 kg of pelleted soyhulls (9.6% CP, DM basis) regardless of RFI rank. Cows had adequate bunk space and ad libitum access to water and mineral supplement for both grazing periods. Intake data, calculated using RPM data, are presented in Table 1. When cows were gestating and not lactating, –RFI cows consumed 80% as much forage as +RFI cows, and when lactating, –RFI cows consumed 89% as much forage as +RFI cows. We currently believe intake differences between lactation stages is due to greater energy intake by +RFI cows being used for milk production whereas –RFI cows must increase intake to meet milk production needs. The result is that efficiency differences between RFI groups narrows. We measured similar results in RFI-phenotyped fall-calving first-calf heifers. Efficient (–RFI) and inefficient (+RFI) animals were contrasted through the winter while they were grazing stockpiled tall fescue. Cow BW at weaning and calf weaning weight were not different, but forage intake by efficient cows was 87% of intake by inefficient cows (Table 2). These differences agreed with other research (Nkrumah et al., 2004, 2006; Kolath et al., 2006; Castro Bulle et al., 2007) and reduction in intake between groups when lactating compared to non-lactating agreed with mice data (Hughes and Pitchford, 2004). Selection for efficiency can reduce DMI an average of 16% without any notable change in condition or BW of the cow. As generations selecting for efficiency are stacked, it is logical to conclude that intake reduction will be even greater. Progeny generated from matings of animals with known efficiency phenotypes will likewise differ in intake. Table 3 presents data of offspring from dams with known RFI phenotypes. Dams were phenotyped as heifers and then bred to bulls of known RFI phenotype. Dams were grouped one standard deviation above (+RFI; inefficient) and below (–RFI; efficient) mean RFI. Intake was reduced 7.4% and feed conversion was improved 10.9% when progeny from efficient dams were compared with progeny of inefficient dams. Sire selection can have similar impact on progeny performance. Two sires were contrasted in Table 4. Sires were mated to cows such that cow RFI was evenly distributed across sires. Although this only contrasts 2 sires, we have measured similar response across a wider array of sire groups. In this example, progeny from sire B consumed 11.7% less DM and had an improved feed conversion efficiency of 19.4%. Economic impact of selecting for efficiency is substantial enough to pull through industry adoption of this technology. The outcome expected would be more efficient use of forage resources. IMPACT OF ENERGY AND AMINO ACID IMBALANCE ON PERFORMANCE Forage digestion by cattle yields short chain fatty acids and microbial protein from ruminal fermentation. Protein not fermented also passes to the small intestine. Amino acid profile of microbial protein is relatively constant across various diets. Protein present in forages is primarily photosynthetic protein, so likewise, is relatively constant across plants in being highly degraded in the rumen and in AA composition. The ME yield from forage is a function of digestible cell wall content as has been demonstrated by numerous experiments relating animal performance to forage digestibility. Because microbial protein production is a function of fermentable energy and microbial AA composition is constant, microbial AA flow post-ruminally should be relative and constant to forage ME. High rumen degradability and similarity of AA profile across plants results in consistency of absorbable AA profile available to the animal when fed a range of forage diets. What would be expected is that total mass of AA would change as forage protein content and availability changed, but AA profile would remain constant. The hypothesis that AA to ME remains constant is important because it allows us to identify an imbalance of AA to energy that occurs across forage diets. Assuming that rumen-degradable protein (RDP) is not limiting microbial growth in the rumen, AA supply is typically limiting maximum growth potential on forage diets. Figure 1 contrasts the energy (ME) allowable gain and absorbable AA allowable gain for backgrounding calves fed orchard grass hay (2.35 Mcal ME and 16% CP). Allowable gains are calculated as potential gain supported by the ME consumed or by the most limiting AA calculated as AA supplied by microbial protein and undegraded forage protein (70% RDP). This hypothetical forage would contain enough energy to support 0.9 kg/d gain with DMI at 2.25% of BW but only yield post-ruminal methionine (most limiting AA) flow capable of supporting 0.3 kg/d gain. The point made is that imbalance between energy and AA supply exists for cattle fed forage diets. Richardson and Hatfield (1978) reported methionine was first-limiting when post-ruminal protein supply was predominantly of microbial origin. Storm and Orskov (1984) concluded that microbial protein was limiting in methionine, lysine, histidine, and arginine. Greenwood and Titgemeyer (2000) likewise measured methionine to be first-limiting in a structural carbohydrate-based diet. A strong body of research exist demonstrating response to post-ruminal AA 126 the level of protein and AA that are limiting. Correcting imbalances in absorbable AA to energy ratio for high quality forage diets provide enough benefit that they are generally profitable. Application spans from backgrounding calves to young cows. supply in cattle fed forage-based diets, indicative of an imbalance between absorbable AA supply and ME availability. Imbalance of AA and ME will be most pronounced in cattle capable of growth (backgrounding calves, developing heifers, pregnant heifers, or young cows). Forcherio et al. (1995) evaluated this concept using lactating two-year old cows grazing spring tall fescue pastures. Each cow was supplemented with 1 kg/d of corn mixed with soybean meal or bloodmeal to provide 54 or 250 g of rumen-undegraded protein (RUP), respectively. Supplementing RUP improved cow ADG, reduced milk production, and had no effect on calf ADG (Table 5). Concluded from this study was that RUP supplementation improved gain by partitioning tissue accretion to the animal and away from milk production. When energy was supplemented in absence of UIP, milk production increased and BW loss occurred. It is our belief that correcting imbalance of AA to energy ratio in forage diets can be beneficial to young cows that are biologically committed to competing functions of body tissue accretion and milk synthesis. We have measured beneficial responses in developing heifers to methionine supplementation. Heifers were fed supplements containing 0, 7.5, or 15 g of MFP (rumen stable methionine) daily for 85 d (Table 6). The higher level of methionine supplementation increased ADG and reproductive tract score of heifers. We concluded that heifers fed supplemental methionine were more physiologically mature due to more rapid growth during development. Rumen escape of MFP is reported to be 40%. Our nutritional model predicted 6 to 7 grams of additional methionine were needed to maximize ADG on the forage diet fed to these heifers. Prediction of additional methionine required to improve gain and improvement in gain from supplying 6 to 7 grams additional methionine were in agreement. Rodriquez et al. (2009) fed stocker steers supplement containing either corn gluten meal or Alimet (rumen stable methionine) to supply sulfur AA post-ruminally. Supplying methionine individually increased gain of pastured cattle similar to providing RUP as corn gluten meal (Table 7). Gain response was also level-dependent for both protein sources. We also provided supplemental RUP in the form of bloodmeal to calves grazing alfalfa pasture. Although alfalfa protein was high, supplemental RUP improved gain (Table 8). Data can be interpreted supporting the conclusion that growing cattle consuming forage diets have an imbalance of absorbable AA supply relative to energy availability. This imbalance is created by discrepancy in microbial AA profile relative to animal AA requirement. It is my present belief that gain improvements occurring from increased supply of AA post-ruminally are due to increased lean tissue accretion at the expense of lipid accretion. When energy supply is greater than absorbable AA supply, energy is deposited as lipid. However, when absorbable AA supply is increased to match energy supply, lean tissue will be accreted. The greater water content of lean tissue results in measurement of improved daily gain. This view is corroborated by response to supplemental RUP having a point of diminishing return. There is general agreement on CONCLUSIONS Technology now being adapted can move beef cows to consuming 20% less forage while maintaining growth and milk production. While it is unknown how efficient animals adjust nutrient requirements or if they even do, we need to consider how potential for reduced intake will affect intake of protein and minerals. It has been demonstrated that imbalance exists between amino acid and energy in cattle fed forage diets. Will this be exacerbated by more efficient cattle that consume less feed? Likewise, as we use and develop growth promotants for forage-fed cattle, the impact of amino acid-energy imbalance needs to be considered. Can a calf grazing high-quality pasture maximally respond to a growth implant if inadequate amino acids are supplied to support lean growth? Production economics likely will continue to select a heavier calf for feedlot production, giving greater profit potential for backgrounding calves. As methods are developed to select for more efficient cattle, forage intake should decrease. Greater value of backgrounding and change in intake potential by cattle emphasizes the importance of correcting nutritional imbalances in cattle fed forage. LITERATURE CITED Castro Bulle, F. C. P., P. V. Paulino, A. C. Sanches, and R. D. Sainz. 2007. Growth, carcass quality, and protein and energy metabolism in beef cattle with different growth potentials and residual feed intakes. J. Anim. Sci. 85: 928-936. Forcherio, J. C., G. E. Catlett, J. A. Paterson, M. S. Kerley, and M. R. Ellersieck. 1995. Supplemental protein and energy for beef cows consuming endophyte-infected tall fescue. J. Anim. Sci. 73:3427-3436. Greenwood, R. H. and E. C. Titgemeyer. 2000. Limiting amino acids for growing Holstein steer limit-fed soybean hull-based diets. J. Anim. Sci. 78:1997-2004. Hughes, T. E. and W. S. Pitchford. 2004. How does pregnancy and lactation affect efficiency of female mice divergently selected for post-weaning net feed intake? Aust. J. Exp. Agric. 44:501-506. Mertens, D. R. 1987. Predicting intake and digestibility using mathematical models of ruminal function. J. Anim. Sci. 64:1548-1558. Meyer, A. M., M. S. Kerley, and R. L. Kallenbach. 2008. The effect of residual feed intake classification on forage intake by grazing beef cows. J. Anim. Sci. 86: 2670-2679. Nkrumah, J. D., J. A. Basarb, M. A. Price, E. K. Okine, A. Ammoura, S. Guercio, C. Hansen, C. Li, B. Benkel, B. Murdoch, and S. S. Moore. 2004. Different measures of energetic efficiency and their phenotypic relationships 127 Richardson, C. R. and E. E. Hatfield. 1978. The limiting amino acids in growing cattle. J. Anim. Sci. 46:740-745. Storm, E. and E. R. Orskov. 1984. The nutritive value of rumen micro-organisms in ruminants. 4. The limiting amino acids of microbial protein in growing sheep determined by a new approach. Br. J. Nutr. 52:613-620. with growth, feed intake, and ultrasound and carcass merit in hybrid cattle. J. Anim. Sci. 82:2451-2459. Nkrumah, J. D., E. K. Okine, G. W. Mathison, K. Schmid, C. Li, J. A. Basarab, M. A. Price, Z. Wang and S. S. Moore. 2006. Relationships of feedlot feed efficiency, performance, and feeding behavior with metabolic rate, methane production, and energy partioning in beef cattle. J. Anim. Sci. 84: 145-153. Table 1. Forage intake by fall-calving cows of different residual feed intake (RFI) phenotype (one-third most and least efficient1) –RFI RFI, kg/d -4.37 Initial BW, kg 591 Initial BCS 5.3 Nonlactating DMI, kg/d 12.4 Lactating (Pairs) DMI,kg/d 12.5 1 Efficient (–RFI) and inefficient (+RFI) cows. +RFI 5.04 565 5.3 % Difference 15.6 79.5 14.1 88.6 Table 2. Forage intake, average cow body weight, and calf weaning weights contrasted between efficient (–RFI) and inefficient (+RFI) cows Forage intake, kg/d Calf weaning weight, kg Cow BW, kg a,b Means with unlike superscripts differ (P < 0.10) –RFI 14.1b 260 616 + RFI 16.2a 273 620 Table 3. Effect of dam phenotypic selection for residual feed intake (RFI) on progeny performance RFI group1 Dam RFI Calf ADG, kg/d Calf intake, kg/d Calf FCR2 Calf RFI - RFI -1.73 1.75 9.55 5.46 -0.41 µ RFI 0.14 1.76 10.02 5.69 0.07 + RIF 1.77 1.68 10.31 6.13 0.29 -RFI vs. +RFI, % + 4.2 -7.4 -10.9 1 Dams were grouped one standard deviation above (+RFI; inefficient) and below (-RFI; efficient) mean RFI. 2 Feed conversion ratio. Table 4. Effect of sire phenotype selection for residual feed intake (RFI1) on progeny performance Calf ADG, kg/d Sire A (+RFI) 1.72 Sire B (-RFI) 1.88 % Change + 9.3 1 Efficient (–RFI) and inefficient (+RFI) sire. 2 Feed conversion ratio. Calf intake, kg 10.79 9.53 -11.7 128 Calf FCR2 6.28 5.06 -19.4 Calf RFI 0.73 -0.60 Table 5. Effect of rumen undegradable protein (RUP) supplementation on two-year-old cow performance RUP1 Tall fescue2 + E+ Ea b Cow ADG, kg/d 0.1 -0.10 -0.2 0.2 Calf ADG, kg/d 0.9 0.9 0.9 1.0 Milk intake, kg/d 7.3 9.8 8.2 9.6 1 Ad libitum access to tall fescue plus 1.07 kg/d of supplement consisting of cracked corn with soybean meal (-) or blood meal (+) to provide 54 or 250 g of RUP, respectively. 2 Free-choice mineral provided ad libitum to endophyte infected (E+) or free (E-) tall fescue pasture. a,b Means with unlike superscripts differ (P < 0.05). Table 6. Effect of methionine (MFP) supplementation on reproductive development of heifers MFP, g/d 7.5 369 0 374 Initial BW, kg ADG, kg/g day 0 to 30 0.8b day 0 to 85 0.8 Pelvic area, cm2 184 1 Reproductive tract score 4.2b 1 1 = infantile, 5 = mature. a,b Means with unlike superscripts differ (P < 0.05). 15 374 1.0a,b 0.8 186 4.2b 1.2a 0.9 183 4.5a Table 7. Effect of rumen undegradable protein (RUP) and rumen-protected methionine (Alimet) on growth performance of stocker calves fed Bermuda grass hay Control TSAA1 from CGM, g/d 2 4 6 Intake, kg/d Hay 4.2 4.6 Supplement 2.0 2.0 ADG, kg/d 0.5 0.6 1 Total sulfur amino acids from corn gluten meal. 4.9 2.0 0.7 4.7 2.0 0.7 2 4 5.1 2.1 0.6 5.0 2.1 0.6 Alimet, g/d 6 5.2 2.1 0.7 8 5.0 1.9 0.6 Table 8. Effect of rumen undegradable protein (RUP) supplementation on daily gain of calves rotationally grazing alfalfa pastures Control Energy1 100g BM2 200g BM3 ADG, kg/d 0.6 0.7 0.8 0.8 1 Heifers were supplemented with ground corn at 0.3% of BW. 2,3 Heifers were supplemented with ground corn at 0.3% of BW and 100 or 200 g of blood meal. 129 1.6 1.4 ADG, kg/d 1.2 1 Energy 0.8 Amino Acid 0.6 0.4 0.2 0 0 1 2 3 4 DMI, % BW Figure 1. Comparison of gain potential based upon 2.35 Mcal ME and 16% CP orchard grass hay at increasing intake (% of BW). 130 POTENTIAL FOR NUTRITIONAL IMBALANCE IN HIGH-QUALITY FORAGES M. Kerley Notes Proceeding, Grazing Livestock Nutrition Conference July 9 and 10, 2010 APPLICATION OF NUTRIENT REQUIREMENT SCHEMES TO GRAZING ANIMALS H. Dove*1, S. R. McLennan†, and D. P. Poppi‡ * † CSIRO Sustainable Agriculture National Research Flagship, G.P.O. Box 1600, Canberra, A.C.T. 2601, Australia Queensland Primary Industries and Fisheries, Animal Research Institute, Yeerongpilly, Qld 4105, Australia; and ‡ Schools of Animal Studies and Veterinary Science, University of Queensland, Gatton, Qld 4343, Australia ABSTRACT: The ability to apply nutrient requirement schemes to grazing animals can be related to the purpose of the application, the extent to which the requirement scheme truly represents animal requirements, the nature of input data required about the plants and the animals, and the nutrient supply that results from the diet composition and intake of the animals. These issues are discussed in relation to the published energy and protein requirements within the American (NRC, 2007) and Australian (CSIRO, 2007) requirement schemes. Data are presented to indicate that with respect to both energy and protein, the differences between schemes are not great, especially when differences in assumptions of the schemes are taken into account. An understanding of these assumptions is thus critical to the field application of published requirements. With respect to protein requirements, it is suggested that further work is needed to incorporate N recycling into published requirement schemes. In housed animals, live weight gains are satisfactorily predicted by both schemes, but it is suggested that the greatest impediment to the field application of nutrient requirements is the difficulty of obtaining data on nutrient supply (i.e., data on the diet composition and intake of grazing livestock). The Australian requirement system, as captured in the decision support tool GrazFeed, incorporates an algorithm to predict actual intakes based on pasture and animal data. Data are presented to indicate that with temperate pastures, this approach allows accurate prediction of live weight gains of sheep and cattle under field conditions, but it is less successful under tropical and rangeland conditions, for which the intake algorithm is not designed. Under tropical conditions, the application emphasis may shift from the prediction of animal performance to the prediction of intake from observed animal performance, as a means of fodder budgeting; ongoing work in this area is discussed. Finally, it is suggested that the field application of nutrient requirement schemes, particularly by producers, will be constrained as long as this involves the manipulation of complex equations and tabulated data. Such application could be accelerated if it were supported by software packages that could be used by producers or their advisors and that could assist the complex calculations and allow them to be conducted for temperate, rangeland and tropical grazing conditions. INTRODUCTION Most of the world‟s domestic ruminants obtain their daily requirements of nutrients from improved pasture, natural pasture, or browse. Any attempt to apply published nutrient requirements under grazing conditions therefore must also consider the amount and quality of the forage resource because as this resource responds to regional climate, seasonal weather conditions, and the plant species present, and these factors influence the amount and quality of the diet selected by the animals. Nutritional management of livestock thus implies an understanding of nutritional supply from plants which, by contrast with housed animals, is difficult to define and ever-changing in terms of both quantity and quality. This means that a key aspect of the application of nutrient requirement schemes becomes a consideration of the accuracy of the published nutrient requirements relative to the accuracy with which nutrient supply can be quantified. Research and discussion on the nutrient requirements of ruminants has continued for more than a century. As Corbett and Freer (2003) emphasized, in an historical context, this has been driven in part by the need in cold climates to house livestock for some of the year, which generates a need for data about diet quality and animal requirements to permit diet formulation for a desired level of animal performance. However, the main issue for the nutritional management of grazing animals is mostly the reverse, of predicting likely performance from a forage intake usually poorly quantified (Corbett and Freer, 2003; Dove, 2009). The nutrient requirements of grazing ruminants have been reviewed and tabulated recently by CSIRO (2007) and also in a series of National Research Council publications (NRC, 2000; 2001; 2007). These publications themselves draw on earlier feeding standards (e.g., NRC, 1981; 1985; INRA, 1989; SCA, 1990; AFRC, 1993; Tamminga et al., 1994). All provide information on the daily requirements of livestock for energy, protein, minerals, vitamins, and water for the purpose of maintenance and the additional requirements for live weight gain, pregnancy, lactation, and fiber production. In this paper, we will restrict our discussion to the requirements for and the supply of energy and protein in grazing ruminants. We do not present a detailed comparison of the various nutrient requirement systems; rather, our emphasis will be on the application of requirements under field conditions. Nevertheless, differences exist between the published requirement Key words: nutrient supply, diet composition, intake, digestibility, protein content, nitrogen recycling 1 Corresponding author: [email protected] 133 accuracy of the output decreases across this continuum (i.e., scientists may be seeking a highly quantitative, accurate answer, whereas livestock managers may be content with „an indication‟). The level of accuracy also interacts with the reason for the application. For example, the prediction of animal performance under field conditions implies that as part of the application of requirements, diet composition and intake are being estimated or predicted to compare with requirements and thereby predict animal performance. Prediction in this manner demands a greater complexity of input and accuracy of output than, for example, backcalculating from observed animal performance and published nutrient requirements to get an indication of forage intake for fodder budgeting purposes (e.g., Baker, 1982; 2004; and see below). A related issue to our question is the ease with which the various requirement systems allow application under field conditions. For example, if the information about requirements is only available as complex equations or as tabulated data (e.g., NRC, 2007), for some users (especially producers) application under field conditions may be more difficult than if the requirement system is supported by a user-friendly, computer-based package that encapsulates the system for field use. An example of the latter is the use of the GrazFeed and GrassGro software packages (Freer et al. 1997; Moore et al., 1997) to support field application of both the SCA (1990) and CSIRO (2007) systems. systems, some of which are discussed in more detail below, but we will argue that these are not the major constraint in applying feeding standards to the grazing animal. More usually, the constraint to the field application of published requirements has been the quantification of the interaction between the animal and the available forage in the determination of the animal‟s dietary intake and the nutritive value of its diet. This constraint led Alderman (1978) to suggest that in the grazing situation, it was too difficult to use published nutrient requirements and that “…emphasis on quantitative nutrition has to be replaced by…expected yield of forages, optimization of stocking density and measurement of animal production as an estimate of feed intake.” An emphasis on forage yield and stocking density in grazing management is appropriate, but as Corbett and Freer (2003) discussed, the abandonment of quantitative nutrition and the use of animal production to infer intake severely constrain the application of feeding standards to predict livestock performance under field conditions, although the latter may be useful for fodderbudgeting purposes (see below). Conceptually, the application of nutrient requirement systems to grazing livestock can be addressed in terms of several key questions: 1. Who is doing the applying and for what purpose? The reason for applying the system and how this is done will determine, in part, the level of accuracy that is acceptable. 2. How well do the various systems estimate key requirements? Are the published requirements sufficiently quantitative and accurate that they represent the transactions in digestion and metabolism and thus serve as a „target‟ against which to balance nutrient supply in the field? 3. What plant and animal inputs are required to apply the systems under grazing conditions; is it possible to obtain this information easily, and are the techniques for defining the nutritive value (NV) of the dietary components sufficiently accurate to allow NV to be related to nutrient requirements? 4. What nutrient supply results from grazing the forage resource? In short, how can we estimate diet composition and intake, and will the system accommodate the provision of extra nutrients as supplementary feed? Points 2, 3, and 4 clearly indicate that, under field conditions, the issues are those of determining whether the requirement systems accurately define the performance of the animals from known nutrient intake, and whether the necessary inputs can be described in terms of the intake and diet quality with sufficient accuracy to quantify this nutrient intake. HOW WELL DO THE DIFFERENT SYSTEMS ESTIMATE REQUIREMENTS FOR THE MAJOR NUTRIENTS? After more than a century of effort to define the nutrient requirements of livestock, our understanding of ruminant digestion and metabolism is sufficiently advanced that most of the published requirement systems are probably accurate enough to permit field application. This is not to say that further refinements are not possible, particularly in relation to the validity of some of the assumptions underpinning the estimation of requirements. As often as not, differences in assumptions explain the major differences between the various schemes; this can be illustrated by a consideration of recently derived energy and protein requirements. Energy Requirements Modern systems for expressing the energy requirements of livestock usually do so either in terms of metabolizable energy (ME; SCA, 1990; AFRC, 1993; CSIRO, 2007), net energy (NE; INRA, 1989), or both (NRC, 1985; NRC, 2007). The efficiency of conversion of ME to NE varies with the purpose for which ME is used and is greatest for maintenance. Moreover, to varying degrees in the different requirement schemes, the efficiencies of conversion of ME to NE for maintenance (km), growth (kg), and lactation (kl) are themselves regarded as functions of dietary quality, conveniently described by M/D (MJ ME/kg DM). It follows that it is not possible to attribute a single NE value to a feed. By contrast, the ME content of a feed can be well predicted from other, simpler measures of information such as its DM or OM digestibility. WHO IS DOING THE APPLYING AND FOR WHAT PURPOSE? Although who and what questions in the section title may seem simple, it is fundamental to the extent that it determines, in part, the level of accuracy that is acceptable to the user. Conceivably, nutrient requirement systems could be applied under field conditions by scientists, by agricultural advisors, and in some cases, by livestock producers themselves. It is likely that the necessary 134 weight. For example, in beef cattle, NRC (2000) used changes in BCS as a means of practically assessing the energy requirements of mature cows. For field application, this might be more practical than a system based on measuring changes in live weight. Despite such issues, as a generalization, the current accuracy of published energy requirements seems sufficient to allow their application in the nutritional management of grazing livestock, at least in temperate regions. There are differences in the published estimates of energy requirements between the different systems. For example, the ME requirement of a 30-kg lamb growing at 200 g/d is about 16% greater in the Australian system (CSIRO, 2007) compared with AFRC (1993). This is due, in part, to the greater maintenance requirement within the former estimate. By contrast, the ME requirements of housed mature, dry (non-lactating) ewes maintaining weight are very similar between the Australian (CSIRO, 2007) and USA (NRC, 2007) systems (Figure 1). For housed ewes consuming equivalent diets, the difference in estimated ME requirements for maintenance is only 2 to 3%, which would be unlikely to be of practical significance. The requirements for grazing ewes using CSIRO (2007) are 17 to 18% greater than those for housed ewes estimated using NRC (2007), reflecting the 15% increment for grazing embedded in the former estimate. A similar comparison for lactating ewes with a single lamb in early lactation highlights the need for an understanding of the assumptions underpinning requirement estimates (Figure 2). In this case, for equivalent conditions, the total ME requirements estimated from CSIRO (2007) are on average about 10% greater than those tabulated in NRC (2007). Investigation of the source of this difference in estimates shows that it is related to two assumptions used in the CSIRO (2007) calculations that are not in the NRC (2007) estimates. The first relates to the respective assumptions concerning kl; in NRC (2007) this is assumed to be constant (0.644), whereas in CSIRO (2007) it is a function of dietary M/D and for the dataset in Table 15.1 of NRC (2007) varies from 0.56 to 0.60. Recalculation of NRC (2007) requirements using a variable kl increases estimated requirements and brings them within 5% of those of CSIRO (2007). The second assumption relates to the maintenance component of the total requirement, which in NRC (2007) is not related to the level of production. By contrast, the CSIRO (2007) estimate of maintenance is, in part, a function of the level of production in the ewe [0.10 * (ME for milk + ME for live weight change)]. Further recalculation of the NRC (2007) estimates by incorporating an effect of level of production on maintenance gives estimated requirements that are statistically similar to those of CSIRO (2007; Figure 3). This example emphasizes the point that although „differences‟ in estimated requirements might be „arithmetically real‟, they are often not large and are often explicable in terms of the assumptions underlying the different estimates. In turn, this indicates the need to appreciate the assumptions involved. With continued research, and as the validity of these various assumptions is investigated further, refinements to published ME requirements of livestock can be expected. This is particularly the case in pregnancy and (or) lactation, during which it is very likely that patterns of true maternal weight loss can be masked by changes in tissue hydration (see CSIRO, 2007). These changes mean that the ME supplied by weight loss during lactation, or the amount of extra grazing or supplement required to compensate for the weight loss, will be underestimated. To an extent, the issues arising from changing tissue hydration can be overcome by monitoring body condition score (BCS) rather than live Protein Requirements Under grazing conditions, the protein content of green feed may often be sufficient to meet the protein requirement of grazing livestock, although grazing, high-producing dairy cows may have a need for extra protein. In animals grazing low-quality roughages, such as dry summer pasture, crop residues, or dead herbage in the dry tropics, the need for protein becomes a major issue. A system for defining protein requirements is thus needed to identify and quantify the need for protein supplements under these circumstances. Such a system must account for: the needs of the rumen microflora for soluble N (degradable protein intake; RDP) and for fermentable energy, to optimize fiber digestion and the production of microbial protein; the need with some productive purposes for an increased supply of dietary AA post-ruminally (undegraded dietary protein; RUP); and feedbacks between tissue AA availability, efficiency of utilization of absorbed AA, and feed intake. Current systems for expressing protein requirements in general accommodate these needs, although to different extents. Ultimately, the usefulness of any requirement system also has to be a balance between the precision with which it describes rumen digestive processes and the difficulty of obtaining the input data required to describe these processes. This is particularly relevant in estimating protein requirements because of the difficulty of defining the degradability of forage protein. For the grazing animal, the lower precision with which the manager can define the nutrient supply is usually a greater constraint than the accuracy of requirements. Protein requirements were expressed in terms of „absorbed protein‟ by NRC (1985) and of „metabolizable protein‟ (MP) by AFRC (1993) and by NRC (2007), although the latter also tabulates requirements in CP terms for different levels of degradable intake protein. The SCA (1990) publication expressed requirements as „apparently digested protein leaving the stomach (ADPLS),‟ whereas the French system (INRA, 1989) used „protein truly digested in the small intestine‟ (PDI). The recent Australian system (CSIRO, 2007) moved from ADPLS to „truly digested protein leaving the stomach‟ (DPLS) by estimating total protein requirements at the tissue level and converting from this to DPLS using an assumed efficiency of conversion of DPLS of 0.7 for all purposes except wool growth, for which efficiency of conversion of DPLS to wool protein was assumed to be 0.6. The DPLS requirements for a growing lamb presented in CSIRO 135 (2007; see their Table 2.5) are less than would be estimated from AFRC (1993). However, for live weight gains of 100 to 250 g/d in a 4-month old, 30-kg lamb (maturity 0.6), the DPLS requirements of 68 to 88 g/d that can be calculated in CSIRO (2007) are similar to the MP requirements of 74 to 89 g/d tabulated in NRC (2007, see their Table 15.2). In making such comparisons between estimated requirements, it is important to appreciate how the magnitude of any differences between requirement systems compares with the likely effects of plant species or stage of growth of the herbage consumed. Yu et al. (2003) compared the Dutch DVE/OEB system (Tamminga et al., 1994; DVE = truly absorbed protein in the small intestine; OEB = degraded protein balance) with the NRC (2001) system, for predicting the supply of protein to dairy cows from alfalfa and timothy cultivars at various stages of growth. The „degraded protein balance‟ (OEB) compares the potential microbial protein synthesis from RDP consumption vs. that from the energy supplied by rumen fermentation. Positive values indicate potential N loss from the rumen, whereas negative values imply a possible N shortage in the rumen; an optimum OEB should thus be close to zero (Yu et al., 2003). These calculations do not take into account recycling of N to the rumen, which can be substantial. When the predicted DVE supply from the various feeds (g/kg DM) was compared, the slope of the line relating NRC (2001) estimates and the DVE/OEB system did not differ from 1, although there was a significant difference in intercepts, indicating a positive bias of 4 to 10 g DVE/kg DM in the NRC (2001) estimates (Figure 4). However, note that the effect of pasture species (timothy vs. alfalfa) or stage of growth on absorbed protein supply was much greater than the difference between the systems. This trend is even more marked in RDP balances predicted by the two systems, which were not different statistically (Figure 5). By contrast, there were marked differences between pasture species and within the species differences, between stages of growth. There were also differences between timothy cultivars, but not between alfalfa cultivars. These data serve as a reminder that a proper evaluation of the forage resource and its NV may be of greater quantitative significance than any concern about possible differences between requirement systems. Nitrogen recycling to the rumen is not accounted for in any system, and in itself, is not of any importance within high-N, temperate pasture systems. However, in the desire to decrease N excretion and its environmental consequences, this issue is being revisited. It would be an advantage to enhance N recycling to the rumen, especially with highproducing dairy cows, so as to decrease the calculated N (or CP) requirement (Reynolds and Kristensen, 2008). This affects the calculation of N or RDP balance across the rumen and the requirements for quantity and type of supplement to be used. The issue is even more critical with low-N content forages. For example, in tropical northern Australia, for a speargrass (Heteropogon contortus) pasture of 50% OM digestibility, 40 g CP/kg DM, and a microbial protein yield of 130 g MCP/kg DOM (or 9 g MCP/MJ fermentable ME), the RDP deficit is 27.5 g RDP/kg grass. A 200-kg weaned steer might eat up to 20 g/kg BW/d or 4 kg/d of pasture, resulting in a requirement to supplement with 55.7 g/d of a urea-ammonium sulphate mix. However, if recycling of N of 10.9 g N/d (Dijkstra et al., 1992) is assumed then the required supplement intake is only 27.9 g/d of a urea-ammonium sulphate mix. Interestingly, from practical field supplement experiments with weaned cattle, the current recommended intake of urea-ammonium sulphate mix to optimize the live weight gain response in northern Australia is 30 g/d (Winks et al., 1970). To be useful in livestock nutritional management, MP or DPLS requirements are converted to daily requirements for CP (g/kg DM), as in both CSIRO (2007) and NRC (2007). However, this can introduce a further difficulty in comparing the various protein requirement schemes because of different assumptions about the effective degradation of the dietary protein (Edg). The Edg is the protein degradability (dg) that will occur at a given rumen outflow rate and is the „link‟ between dietary CP content and MP/DPLS. The Edg is difficult to measure in vivo and more frequently is derived from dg, measured as the disappearance of N from a feed suspended in polyester or nylon bags in the rumen for fixed times. It is crucial that the estimate of dg makes allowance for microbial colonization of the sample during rumen incubation, especially for lowquality forages, and the conversion from dg to Edg clearly depends on the assumptions made about the likely outflow rate from the rumen (see CSIRO, 2007 for more detailed discussion). As with energy requirements, continuing refinements of protein requirements can also be expected. For example, requirements may be increased to allow for the interaction between increased protein supply and the capacity of animals to deal with gastrointestinal parasites in both the short and long term (Datta et al., 1999) and especially in late pregnancy (Robinson, 2002). Despite the difficulties in estimating Edg and the need to resolve issues, such as nutrient-parasite interactions, the above discussion suggests that inaccuracies or differences between feeding systems in estimated protein requirements are unlikely to be the main constraint to the application of nutrient requirement data under field conditions. Prediction of Animal Performance Given that, after allowance for differences in underlying assumptions, there is a general similarity between the predictions of different requirement systems, the next crucial question is how predicted requirements relate to observed performance of animals given diets of known intake and nutritive value. We investigated this question by using data from a series of our own pen-feeding experiments (McLennan, 1997; 2004; Bolam, 1998; Marsetyo, 2003) where young, growing Bos indicus crossbred beef steers (n = 240; ca. 12 to 15 months; range 156 to 243 kg live weight) were restricted to pens and fed diets based on generally low-quality tropical forages with a wide range of supplement types and intakes of varying energy and protein content. We compared measured live weight gains (range -0.1 to 1.2 kg/d) with those predicted using either the SCA (1990) equations or the Cornell Net Carbohydrate and Protein System (CNCPS; Fox et al. 2004). The two systems are quite different in their approach 136 to apportioning energy utilization; SCA (1990) and its successor CSIRO (2007), use the ME system, whereas the CNCPS is aligned to the NE system and uses either a simple theoretical model to estimate a value for TDN content (Level 1 approach) of the diet or a mechanistic approach based on pool sizes of carbohydrate and protein fractions and their digestion and passage rates (Level 2 approach). Because various diet descriptors required to run CNCPS level 2 were not available, surrogate values were taken from feed sources of similar chemical composition in the tropical feed library included in the software. Both SCA and CNCPS require an estimate of the mature size of the animal and in both cases, we assumed values suitable for the type of animal used in these trials. Both systems predict live weight gain based on an estimate of energy retention, as determined by various attributes mainly associated with the diet and animal descriptions. With the CNCPS both an ME-allowable and an MP-allowable gain are predicted, and in our simulations, the lesser of these was used as the predicted gain. The SCA equations represented an MEallowable gain, as the effects of MP supply were already accounted for in determining ad libitum intake by the steers. The regressions for these model simulations were as follows: sacrifice in the precision of the predictions. With SCA, the intercept of the regression line was not different from zero (P > 0.05), indicating good prediction at low growth rate, for instance when forages were provided in the absence of supplement. However, the slope was greater than 1 (P < 0.05), showing a trend for increased under-prediction as growth rate increased in response to increased supplement intake. An overall assessment would be that both systems provided a reasonable prediction of growth rate and thus apparently also of the use of energy for maintenance and growth. The prediction of growth rate, as shown above, entails a number of key steps in the requirement systems, including a description of the animal relative to a mature “finished” animal, and determinations of the M/D of the diet, the maintenance requirements of the animal (ME m), and the efficiencies of conversion of ME to NE for various productive purposes (i.e., km and kg, and of the conversion rate from energy retained to live weight gain). Given the small differences in prediction of growth rate by the SCA and CNCPS systems above, it might be tempting to think that the calculated values for the various components of energy use are also similar for the different systems. However, taking one of these elements, ME m, and applying the appropriate equations in each system, it becomes obvious that there are marked differences between systems in the approach taken, assumptions made, and eventual calculated values. With the SCA system, the maintenance requirement increases with increasing energy intake because the equation (Equation 1.22) to calculate it includes a multiple of ME intake (MEI) to account for the higher maintenance requirements of productive animals. The equation, as applicable to a Bos indicus steer is as follows: CNCPS (Level 2): Observed LWG = 0.123 + 0.947 Predicted LWG, (R2 = 0.71; residual standard deviation (rsd) = 0.159; standard error of prediction (sep) = 0.185; bias = 17.3%); SCA: Observed LWG = -0.031 + 1.171 Predicted LWG, (R2 = 0.79; rsd = 0.141; sep = 0.153; bias = 9.6%). Based on these simulations, errors in the application of these systems would arise from two main sources: the general slight under-prediction of growth rate by both of the above equations and the not insignificant variation in the relationships between predicted and observed live weight gains (i.e., R2 = 0.71 or 0.79 with rsd values of between 0.14 and 0.16 kg/d and sep values from 0.130 to 0.185 kg/d). Such variability is not unexpected with biological data of this nature and the extent to which it can be tolerated depends on the purpose for which the predictions are being made, as alluded to earlier. With the CNCPS, the slope of the trend line was not different from 1 (P > 0.05), but the intercept was greater than zero (P < 0.05), indicating a constant under-prediction of growth rate (approximately 0.08 kg/d) over the full range. When the CNCPS Level 1 approach was taken, the predictions were comparable with CNCPS Level 2: Observed LWG = 0.148 + 0.894 Predicted LWG, (R2 = 0.77; rsd = 0.147; sep = 0.179; bias = 22.3%). There was a similar slight under-prediction of performance at low growth rates using Level 1, with the intercept differing from zero (P < 0.05), but a tendency for over-prediction as growth rate increased, although the slope of the trend line was not different from 1 (P > 0.05). From a practical point of view this is an important finding, as it suggests that for workers in the field with limited information about the composition of the diet of grazing animals, or with limited experience in the use of CNCPS, the simpler Level 1 approach can be used without major MEm = 0.26W0.75exp(-0.03A)/km + 0.09 MEI, where A is age in years. By contrast, there is no adjustment for energy intake in the determination of maintenance requirements in the CNCPS system. In this case, NE for maintenance (NEm) is calculated as a function of metabolic body size with adjustments for breed, physiological state, activity, urea excretion, acclimatization, and heat or cold stress. In our simulations, the main variable between treatments was the weight of the steers, and when NEm was expressed on a BW basis, it was relatively constant as energy intake increased. However, computed MEm on a live-weight basis decreased with increasing energy intake; accordingly km increased. The regression equations for changes in estimated ME m with increasing energy intake are as follows. SCA: MEm = 476.4 + 0.049 MEI, (R2 = 0.52; rsd = 6.91) CNCPS: MEm =514.2 – 0.032 MEI, (R2 = 0.06; rsd = 18.36), where both MEm and MEI are expressed as kJ/kg W0.75.d-1. The values for MEI were determined using the observed intakes by the cattle but the separate calculations of the energy density of the diet (MJ/kg DM) for each system. The low R2 value for CNCPS can be attributed to outliers, which were associated with animals with high intakes of 137 development and use of fecal near-infra-red reflectance spectroscopy (F.NIRS; Lyons and Stuth, 1992; Coleman and Henry, 2002; Dixon and Coates, 2005; 2009) for estimating the quality of the diet selected by grazing animals. The F.NIRS approach relates dietary attributes, such as crude protein content and DM digestibility (DMD), to the NIR spectra of the feces collected from the animals. This allows rapid, frequent, and relatively inexpensive determination of diet quality that can be used as an input for the requirement systems to estimate herbage intake and NV, which ultimately can be used for prediction of animal performance. In improved temperate pastures, once the digestibility of herbage falls below about 70%, there is a marked and increasing reduction in herbage intake. The effect on live weight gain in young animals will be proportionately more marked because the efficiencies of utilization of ME for both maintenance and gain also decreased as digestibility decreases (see CSIRO, 2007). Dietary ME content is difficult to measure directly and is usually obtained by indirect calculation and expressed as M/D = MJ ME/kg DM. This is a key „driver‟ of the estimation of requirements in both the CSIRO (2007) and NRC (2007) systems, and the prediction of intake in the former. As a result, the output from the application of the system under field conditions is sensitive to the M/D value used, and considerable research effort has gone into M/D prediction from DM digestibility (DMD). For example, CSIRO (2007) uses the following equations to calculate ME content of roughage or concentrate feeds in terms of M/D, from DMD (%): cottonseed meal (up to 2%W/d), where maintenance requirements were increased to account for the ME cost of urea excretion. Similar differences between systems could be shown for the other elements of energy use, but in combination, the predictions of animal performance may be quite similar as shown above. This fact highlights the importance of using the system in totality rather than selecting different elements from different systems in isolation. WHAT PLANT AND ANIMAL INPUTS ARE REQUIRED TO APPLY NUTRIENT REQUIREMENT SYSTEMS UNDER GRAZING CONDITIONS? The above question actually comprises two parts. First, can we adequately describe the livestock themselves? Second, can we define the NV of the diet available to or consumed by the grazing livestock? The first question is far easier to answer than the second. As discussed in more detail below and as emphasized by Dove (2009), quantifying what grazing livestock eat, and how much, remains the major constraint to applying nutrient requirement data under field conditions (esp. rangeland conditions). In describing the livestock themselves, their breed, sex, physiological state, and current weight are of obvious importance, although experience in Australia with the field application of nutrient requirements demonstrated that current weight was not enough. Some allowance was required for differences in the mature size of different sheep and cattle breeds, otherwise the same nutrient requirement or herbage intake would be predicted for a large, lean animal and a small, fat animal of the same current weight. SCA (1990) and subsequently CSIRO (2007) addressed this issue by scaling many of their functions to a „standard reference weight‟ (SRW), defined as the weight of the mature female in the middle of the condition score range, for each breed of sheep and cattle. This approach has simplified these systems by allowing generalized equations for sheep and cattle, in which only the coefficients change with species. It has become a cornerstone concept of these systems, but because it influences so many equations in the systems, it is also most important to use an appropriate SRW when applying the systems. In general, defining SRW has been more difficult for users than defining current weight. The main drivers of herbage intake (and thus nutrient supply) are the amount of herbage and its NV; the two most important NV measurements required are the digestibility and CP content. Digestibility is not routinely measured in vivo, but rather estimated indirectly using techniques such as an in vitro procedure. Obtaining an accurate estimate of diet quality is problematical under grazing conditions due to the difficulty of mimicking diet selection by the animal by sampling of the pasture on offer. Grazing animals invariably select a higher-quality diet than the average of or that sampled from the pasture, and the quality of this diet is constantly changing with changing pasture conditions. One of the encouraging developments in this area in the southern USA and northern Australia over the last decade has been the Roughage M/D = 0.172 DMD – 1.707; Concentrate M/D = 0.134 DMD + 0.235 EE + 1.23. The prediction of M/D for concentrates is markedly improved if ether extract (EE, %) is included in the equation, but its inclusion for roughages adds very little to the precision because of the low EE levels in roughage (usually well under 5%; see CSIRO, 2007). For a more complete discussion of the factors influencing predicted M/D of feeds, the reader is referred to CSIRO (2007). One of the difficulties often confronting users of the requirement systems is that of knowing which equation to use to calculate M/D. Even small differences in the calculated value can have marked effects on predicted animal performance by virtue of the fact that it is the major determinant of the efficiencies ME use and thereby also contributes to the calculation of maintenance requirements. Similarly, Edg can also be calculated from the in vitro DMD (%) of the plant material. For example, the following equations (Equations 2.13 and 2.14 of CSIRO 2007, with rumen fractional outflow rate = 0.02), provide estimates of Edg corrected for microbial colonization of the sample, for temperate and tropical forages, respectively: Edg = 0.1852 + 0.01 IVDMD; Edg = 0.408 + 0.0072 IVDMD. In general, the NV of a known ruminant diet can thus be described in terms that allow a comparison with published 138 The provision of pasture with high legume content is a potent way for the manager to ensure greater nutrient intakes in livestock, because they will consume 15 to 20% more legume forage than grass forage of the same digestibility (Freer and Jones 1984). The effect on live weight gain is proportionately greater because the products of legume digestion are also used with greater efficiency for growth. Nonetheless, the evidence that grazing sheep actively select for legume is equivocal. This aspect is discussed in detail by NRC (2007, see Chapter 3). Although there is some evidence of selection for white clover, this does not seem to be the case with subterranean clover in Australia. As a result, in its intake prediction algorithms, the CSIRO (2007) system incorporates effects of herbage mass or herbage digestibility, but does not explicitly model selection for legume. If pasture legume content is increased, predicted intake will increase (Freer and Jones, 1984), but the intake will contain the same legume content as the sward. The relative importance of herbage supply and sward legume content in managing intake depends on how much herbage is present. At low herbage masses (e.g., 0.5 to 1.0 t DM/ha in a sheep system), an increase in sward legume content will usually result in a smaller increase in live weight gain than an increase in herbage mass, which is the greater constraint to intake. In short, emphasis on sward legume content when very little herbage is available is misplaced management; it is more important to increase the quantity of herbage available. However, as the amount of herbage increases further (e.g., to 1.5 to 2.0 t DM/ha and beyond in sheep systems), at a given legume content, the effect of herbage mass on live weight gain diminishes markedly, as animals rapidly approach their potential intakes (Freer et al., 1997). By contrast, an increase in legume content results in continued increases in live weight gain because the animals can eat more of the legume (Freer and Jones, 1984). Hence, as herbage supply increases, it becomes increasingly important to shift the managerial emphasis to increasing sward legume content. In practice, herbage mass and legume content are not entirely separate and the manager would hope to provide more herbage containing more legume. In an attempt to increase the practical application of their published requirements in grazing livestock, a number of nutrient requirement systems have provided functions for estimating the potential intake of grazing animals; the various approaches are discussed in more detail in NRC (2007, see Chapter 3). However, a major criticism of many systems is that they make no allowance for differences in the mature size of different livestock breeds; predicted intakes would thus be the same intake for a large, lean animal and a small, fat animal of the same current weight. As noted previously, SCA (1990) and subsequently CSIRO (2007) addressed this issue by scaling potential intake functions to SRW, an approach later adopted in part by NRC (2007) in developing functions for potential intake. For example, if Z is the ratio of the current weight of an immature animal to its SRW (both in average condition score), potential intake (I, % of mature weight) of an allgrass diet of non-limiting digestibility is given by: I = 6.8Z – 4Z2 requirements (see CSIRO 2007, NRC 2007 for discussion). However, the quantification of what the grazing animal actually consumes is a more complex problem. Although there are general relationships between intake and various characteristics of the forage on offer, it remains difficult to specify the diet of the grazing animal with the same accuracy as current feeding systems define its nutrient requirements. ESTIMATING NUTRIENT SUPPLY IN THE GRAZING ANIMAL There can be little doubt that the major constraint to the application of published nutrient requirements to grazing animals is estimating the quantity and the NV of the diet actually consumed. The problems associated with these estimates, and the various approaches to such measurements are discussed by CSIRO (2007), NRC (2007) and Dove (2010). No matter how accurate systems for estimating nutrient requirements might become, it remains difficult to compare them with the „nutrient supply‟ side of the balance because diet composition and forage intake are difficult to quantify and continually changing as pasture conditions change. Herbage intake, the key „driver‟ of nutrient supply and for improved pastures, is itself highly responsive to: the amount of pasture present; its digestibility, as affected by season and by species content; and its legume content. These same drivers of intake also apply in native (i.e., unsown) pastures, but as the herbage association tends towards rangeland, the species composition of the plant biomass and its spatial distribution become increasingly important (see Hobbs, 1999 for discussion). As described above, studies with temperate pastures have shown that as the digestibility of the herbage on offer decreases, so too will herbage intake, especially below a digestibility of 70%. It follows that in the nutritional management of livestock grazing temperate pastures, there is little point in worrying about the accuracy of published nutrient requirements if the forage resource is of low digestibility. Managing the pasture to increase forage digestibility and thus intake should be the greater concern, although maintaining RDP supply is also a concern when forage digestibility is low. Under rangeland or tropical conditions, the situation is rather different. The DM or OM digestibility of the plant biomass is frequently less than 70% and the management of protein supplementation is also a major concern when the plant biomass is of low digestibility and protein content. Herbage mass is also a key driver of intake and live weight gain. For example, in a improved pasture of 75% digestibility, there will be little increase in herbage intake by sheep once the amount of herbage exceeds 2 t DM/ha because animals have reached their potential intakes (Freer et al., 1997). Herbage intake in sheep begins to decrease at herbage masses below about 1.2 t DM/ha and declines markedly below 1 t DM/ha. It can be calculated (Freer et al., 1997) that at a herbage mass of 0.5 t DM/ha, predicted intake is only 67% (and predicted live weight gain only 52%) of that at 2 t DM/ha. 139 systems. Observed live weight gains agreed closely with predicted live weight gains (Figure 7) suggesting that both the requirement predictions and the intake algorithms within the package had acceptable accuracy. For housed animals, application of the CSIRO (2007) requirements can be achieved by setting pasture mass to zero, but is also supported by the spreadsheet programs „ME_Required‟ and „CP_Required‟, which are available at the above website. In grazing systems based on rangeland or on tropical pastures, a major challenge that remains is to devise similar algorithms to those described above to allow the routine prediction of intake for given conditions of the plant biomass on offer. However, in many such situations, the information of more value to field workers may not be the prediction of intake and thus growth rate of the animals but a determination of their intake from known growth rate. This could be of great value for fodder budgeting purposes or to set appropriate stocking rates; for example, at the end of the growing season in tropical grazing systems when the available pasture usually represents the available reserves for the ensuing dry season period. If the equations describing the utilization of energy by animals are sound and their application results in relatively accurate prediction of growth rate from known intake and diet composition, or for growth rates in the field (Figures 6 and 7), then it follows that this process can be reversed and intake could be estimated by back-calculation from live weight change and an estimate of the energy density of the diet (M/D). In practice, the live weight change used in the system could be either a measured live weight change or alternatively a prediction based on previous experience in the same area (historical value). This approach to intake prediction has been suggested before, (e.g., MAFF, 1975; Baker, 1982; 2004; Minson and McDonald, 1987). What has usually been lacking is the validation of the predictions under practical feeding conditions. With some indication of the validity of the energy utilization equations in SCA (1990), as described earlier, we have developed our own spreadsheet intake calculator (QuikIntake; McLennan and Poppi, unpublished data), which is currently being validated. The main inputs required for an unsupplemented, confined (no grazing) animal include a description of the animal in terms of its sex, age, current live weight, SRW and its live weight gain (either measured or forecast), and an estimate of the M/D of the diet, which, under grazing conditions, could be calculated from the DMD derived from F.NIRS analysis. The ME intake required to achieve the live weight gain can then be determined and thereby also DMI of the forage by dividing by M/D. Early results indicate some problems with intake prediction are emerging with low-quality tropical forages in the variability of the relationship between predicted and observed DMI for a measured live weight change, which relate to the variability around the equations for energy use (see earlier). For example, for weaned steers consuming a hay of Rhodes grass (Chloris gayana) the predicted compared with observed intake for the measured live weight change was 1.50 vs. 1.67 %W/d, whereas for speargrass it was 1.84 vs. 1.15 %W/d. Thus, although there is confidence in the nature of the equations, there is sufficient variability in predicted intakes to be of concern in A young ewe weighing 0.85 of its mature weight of 80 kg thus has a potential intake of 2.31 kg DM/d. Note that this equation allows for the slight decrease in intake in animals with Z > 0.85 (see NRC, 2007). Functions are also available to adjust potential intakes for physiological state (e.g., lactation) and for diet quality (see CSIRO, 2007 and NRC, 2007 for further details). Only the Australian nutrient requirement system (SCA, 1990; CSIRO, 2007) has attempted the next step of moving from potential intake to an estimate of the actual intake as constrained by prevailing pasture conditions. It does so not by tabulating potential or actual intakes as in, for example, (NRC, 2007) but by capturing the estimated nutrient requirements within decision-support tools that also feature algorithms for predicting herbage intake for given animal and pasture conditions (GrazFeed - Freer et al., 1997; GrassGro – Moore et al., 1997; Freer et al., 2009 and see www.pi.csiro.au/grazplan). The algorithms are based on published relationships between herbage intake and variables, such as herbage mass, herbage height, digestibility, legume content, and in pastures of low digestibility, herbage N content as it drives RDP supply. The average digestibility of herbage, an input provided by the user, is allocated by the algorithm into conceptual „digestibility pools‟ within the herbage, from which the animal progressively selects until its intake limit is reached. It should be noted that in rangeland situations or with tropical pastures, the algorithms may not predict intake accurately because they are derived from relationships established with temperate pastures (Freer et al., 1997; Freer et al., 2009). Nevertheless, the incorporation of the intake-prediction algorithm has had 3 major consequences: 1. It has provided the „technical support‟ and the mathematical power to permit the extensive use of the SCA (1990) or CSIRO (2007) nutrient requirements under normal grazing conditions, including use by graziers themselves. 2. It allows for interactions between pasture and supplement intakes and thus has proved very useful in predicting the need for and the response to supplementary feeds. 3. Indirectly, it has resulted in a population of users who have become trained in the skills of assessing pasture quantity and quality and also animal condition. For example, the use of the GrazFeed package in south eastern Australia increased markedly once formal training in these assessments was offered by Departments of Agriculture. The ability of the CSIRO (2007) requirement system, coupled with the intake-prediction algorithm in the above packages, to predict accurately the live weight gain of animals grazing temperate grasslands can be assessed from Figures 6 and 7. In the former, the live weight gains of lambs grazing low-quality summer pasture with or without supplements of 2:1 oat grain:decorticated sunflower meal (Freer et al., 1985; Freer et al., 1988) are compared with live weight gains predicted from the GrazFeed package. The relationship between observed and predicted gains did not differ from the line of equality. Cohen et al. (2003) used the GrassGro package to make a similar comparison for steers grazing mixed grass-alfalfa pasture in western Canada under a range of management 140 3. Further amendments to existing nutrient requirement estimates will undoubtedly be required to deal with, for example: the interaction in pregnancy between possible increases in tissue hydration with simultaneous maternal live weight loss; the interaction between gastrointestinal parasites and nutrient requirements, especially for protein; the effects of pregnancy and (or) lactation on digesta kinetics and especially possible increases in the efficiency of microbial protein production in late pregnancy; and the possible need for extra dietary protein to support mammary development in late pregnancy. general application. Currently, the application of QuikIntake based on the standard equations seems to generally over-estimate intake for a live weight change or loss around maintenance for low-CP forages. The cause of this variability is still being explored, but is most likely related to different efficiencies of MCP production and hence MP supply, possibly as a function of variable N recycling under these conditions. Extension of this process of intake prediction to the grazing situation presents further challenges. In effect, allowance can be made for the additional energy costs of grazing and foraging of the grazing animal, as well as any cost for temperature regulation under extreme conditions. A simplified calculation of the energy expended in eating and walking has been provided by Freer et al. (2009). However, where there already appear to be inaccuracies associated with energy use for pen-fed animals, adding the extra dimension of predicting energy use for grazing activity is likely to further reduce the precision of intake prediction. When Lardy et al. (2004) applied the NRC beef cattle requirement model (NRC, 1996) to cows grazing rangeland and sub-irrigated pastures in Nebraska, they found that at times predicted energy deficits were excessive and “biologically unreasonable”. Prediction errors were likely to have originated from low predictions of the energy density of the diet selected and the resulting low predicted DMI. This reinforces the point that models predicting grazing animal performance need first to provide a relatively accurate estimate of voluntary intake by the animals before then predicting the use of that ingested energy for production. The multiplication of errors for both aspects, as suggested by our results and those of Lardy et al. (2004), could provide wide discrepancies in predicted vs. observed performance. CONCLUDING REMARKS Current approaches to evaluating the NV of feeds or the nutrient requirements of sheep are sufficiently accurate for livestock managers to use them as the basis for matching nutrient requirement and supply in grazing systems. Quantifying nutrient supply is a much greater challenge; in the future, more use could be made of computer-based decision support tools to address this problem. In attempting to match nutrient supply and requirement in grazing animals, there is a temptation to do so on an individual animal basis because this is how requirements are usually presented. However, in many grazing systems, profit is driven by animal production/ha, not production/animal. From the system-wide standpoint of production efficiency, it is thus often more profitable to schedule key events (e.g., lambing or calving) or increase stocking rates (animals/ha) to the point that for some of the year at least, the intake of individual animals is constrained. For example, lambing well before the spring flush of pasture growth may place pregnant or lactating ewes under nutritional stress, but this approach ensures that their lambs are weaned onto good-quality pasture. Similarly, over dry summers, adult animals may be deliberately managed such that they lose weight to minimize the costs of supplementary feeding, a major discretionary cost for producers. Part of the manager‟s role is to decide whether these intentional short-term imbalances between nutrient supply and nutrient requirement can be tolerated or whether they need to be offset with supplementary feed. Increasingly, a component of this assessment will have to be the perceived welfare needs of individual animals or groups of animals. However, as a general approach it is unlikely that profit from a livestock enterprise will be maximized if the balance between nutrient supply and nutrient requirement is pursued at the individual-animal level. FUTURE ISSUES AND CONCERNS The application of nutrient requirement systems under field conditions has raised a number of concerns about issues that might limit future such applications unless they are resolved. 1. For as long as the user of nutrient requirement schemes is required to work with only complex published equations, plus tabulated data that cover only a subset of the possible field situations, the application of such schemes under field conditions will be constrained. Therefore, with many systems, more support needs to be provided to users to facilitate such applications. The extent of uptake and use of the GrazFeed and GrassGro packages, which encapsulate the SCA (1990) and now the CSIRO (2007) systems, indicates that this support is seen by users as a major issue. 2. Possible application of nutrient requirement schemes in tropical or rangeland environments, with plant associations that are very diverse nutritionally and spatially, is a major concern and should be a focus of further research. The use of F.NIRS approaches to estimate the species composition and (or) the NV of the consumed diet, and possibly its intake, may provide a promising approach to resolving this. Further research is needed to provide the relationships that can be used to build intake-prediction algorithms for these environments. LITERATURE CITED AFRC. 1993. Energy and Protein Requirements of Ruminants. CAB International, Wallingford, UK. Alderman, G. 1978. A practical feeding system for ruminants. Proc. Aust. Soc. Anim. Prod. 12:47-57. Baker, R. D. 1982. Estimating herbage intake from animal performance. Pages 77-93 In: Herbage Intake Handbook. J.D. Leaver (ed) Br. Grassland Soc. Hurley, UK. 141 Baker, R. D. 2004. Estimating herbage intake from animal performance. Pages 95-120 In: Herbage Intake Handbook. P.D. Penning (ed) Br. Grassland Soc. Hurley, UK. Bolam M. J. 1998. Manipulation of the supply of protein and energy to ruminants consuming tropical forage through supplementation strategies. Ph.D. Thesis. The Univ. of Queensland, Brisbane, Queensland 4072, Australia. Cohen, R. D. H., J. P. Stevens, A. D. Moore, and J. R. Donnelly, 2003. Validating and using the GrassGro decision support tool for a mixed grass/alfalfa pasture in western Canada. Can. J. Anim. Sci. 83:171-182. Coleman, S. W. and D. A. Henry. 2002. Nutritive value of herbage. Pages 1-26 In: Sheep Nutrition. M. Freer and H. Dove (eds) CAB Interntl. Wallingford, UK. Corbett, J. L. and M. Freer. 2003. Past and present definitions of the energy and protein requirements of ruminants. Asian-Aust. J. Anim. Sci. 16:609-624. CSIRO. 2007. Nutrient Requirements of Domesticated Ruminants. CSIRO Publishing, Melbourne. Datta, F. V., J. V. Nolan, J. B. Rowe, G. D. Gray, and B. J. Crook. 1999. Long-term effects of short-term provision of protein-enriched diets on resistance to nematode infection, and live-weight gain and wool growth in sheep. Int. J. Parasitol. 29:479–488. Dijkstra, J., H. D. StC., Neal, D. E. Beever, and J. France. 1992. Simulation of nutrient digestion, absorption and outflow in the rumen: model description. J. Nutr. 122:2239-2256. Dixon, R. M. and D. B. Coates. 2005. The use of faecal NIRS to improve nutritional management of cattle in northern Australia. Rec. Adv. Anim. Nutr. Aust. 15:6575. Dixon, R. M. and D. B. Coates. 2009. Review. Near infrared spectroscopy of faeces to evaluate the nutrition and physiology of herbivores. J. Near Infrared Spectrosc. 17:1-31. Dove, H. 2009. Balancing nutrient supply and nutrient requirements in the grazing sheep. Small Rum. Res. 90:(in press). Dove, H. 2010. Keynote: Assessment of intake and diet composition of grazing livestock. Pages 31–54 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Fox, D. G., L.O.Tedeschi, T. P. Tylutki, J. B. Russell, M. E. Van Amburgh, L. E. Chase, A. N. Pell, and T. R. Overton. 2004. The Cornell Net Carbohydrate and Protein System model for evaluating herd nutrition and nutrient excretion. Anim. Feed Sci. Tech. 112:29-78. Freer, M. and J. B. Jones. 1984. Feeding value of subterranean clover, lucerne, phalaris and Wimmera ryegrass for lambs. Aust. J. Exp. Agric. Anim. Husb. 24:156-164. Freer, M., A. D. Moore, and J. R. Donnelly. 1997. GRAZPLAN: decision support systems for Australian grazing enterprises. II. The animal biology model for feed intake, production and reproduction and the GrazFeed DSS. Agric. Systems 54:77-126. Freer, M., H. Dove, A. Axelsen, J. R. Donnelly, and G. T. McKinney. 1985. Responses to supplements by weaned lambs grazing mature pasture or eating hay in yards. Aust. J. Exp. Agric. 25:289-297. Freer, M., H. Dove, A. Axelsen, and J. R. Donnelly. 1988. Responses to supplements by weaned lambs when grazing mature pasture or eating hay cut from the same pasture. J. Agric. Sci. (Camb.) 110:661-667. Freer, M., A. D. Moore, and J. R. Donnelly. 2009. The GRAZPLAN animal biology model for sheep and cattle and the GrazFeed decision support tool. CSIRO Plant Industry Techn. Paper (revised April 2009). At http://www.pi.csiro.au/grazplan/files/TechPaperApr09.p df (confirmed October 28, 2009). Hobbs, N. T. 1999. Responses of large herbivores to spatial heterogeneity in ecosystems. Pages 97-129 In: Nutritional Ecology of Herbivores. H-J.G. Jung and G.C. Fahey (eds) Amer. Soc. Anim. Sci. Savoy, Illinois. INRA. 1989. Ruminant Nutrition: Recommended Allowances and Feed Tables. Institut National de la Recherche Agronomique, Paris. Lardy, G. P., D. C. Adams, T. J. Klopfenstein, and H. H. Patterson. 2004. Building beef cow nutritional programs with the 1996 NRC beef cattle requirements model. J. Anim. Sci. 82:E83-E92. Lyons, R. K and J. W. Stuth. 1992. Fecal NIRS equations for predicting diet quality of free-ranging cattle. J. Range Manage. 45:238-244. MAFF. 1975. Energy Allowances and Feeding Systems for Ruminants. Ministry of Agriculture, Fisheries and Food Tech. Bull. 33. HMSO, London. Marsetyo. 2003. Feeding strategies to reduce intake substitution of forages by supplements in beef cattle. Ph.D. Thesis. The Univ. of Queensland, Brisbane, Queensland 4072, Australia. McLennan S. R. 1997. Developing profitable strategies for increasing growth rates of cattle grazing tropical pastures. Final Report to Meat Research Corporation – Project DAQ.100. 117 pp. McLennan S. R. 2004. More effective supplements for the northern beef industry. Final Report to Meat and Livestock Australia Ltd - Project NAP3.122. 96 pp. Minson, D. J. and C. K. McDonald. 1987. Estimating forage intake from the growth of beef cattle. Trop. Grasslands 21:116-122. Moore, A. D., J. R. Donnelly, and M. Freer. 1997. GRAZPLAN: decision support systems for Australian grazing enterprises. III. Pasture growth and soil moisture submodels and the GrassGro DSS. Agric. Systems 55:535-582. NRC. 1981. Nutrient Requirements of Goats: Angora, Dairy and Meat Goats in Temperate and Tropical Countries. Natl. Acad. Press, Washington DC. NRC. 1985. Nutrient Requirements of Sheep 6th Rev. Ed. Natl. Acad. Press. Washington, DC. NRC. 2000. Nutrient Requirements of Beef Cattle 7th Rev. Ed. Natl. Acad. Press. Washington, DC. NRC. 2001. Nutrient Requirements of Dairy Cattle 7th Rev. Ed. Natl. Acad. Press. Washington, DC. 142 NRC. 2007. Nutrient Requirements of Small Ruminants: Sheep, Goats, Cervids, and New World Camelids. Natl. Acad. Press. Washington, DC. Reynolds, C. K. and N. B. Kristensen. 2008. Nitrogen recycling through the gut and the nitrogen economy of ruminants: an asynchronous symbiosis. J. Anim. Sci. 86:E293-E305. Robinson, J. J. 2002. Review of nutritional standards for sheep. Report for the British Society of Animal Science, 16 pp. SCA. 1990. Feeding Standards for Australian Livestock. Ruminants. Standing Committee for Agriculture and CSIRO, Melbourne. Tamminga, S., W. M. Van Straalen, A. P. J. Subnel, R. G. M. Meijer, A. Steg, C. J. G. Wever, and M. C. Block. 1994. The Dutch protein evaluation system: the DVE/OEB system. Livest. Prod. Sci. 40:139-155. Winks, L., G. I. Alexander, and D. Lynch. 1970. Urea supplements for grazing beef weaners. Proc. Aust. Soc. Anim. Prod. 8:34-38. Yu, P., D. A. Christensen, and J. J. McKinnon. 2003. Comparison of the National Research Council-2001 model with the Dutch system (DVE/OEB) in the prediction of nutrient supply to dairy cows from forages. J. Dairy Sci. 86:2178-2192. ME requirement (MJ/day) 20 17.5 15 12.5 10 7.5 5 30 50 70 90 110 130 150 Ewe mature weight (kg) Figure 1. Comparison of the ME requirements for weight maintenance in housed mature dry ewes of mature weight 40 to 140 kg. NRC (2007) data (■) from their Table 15.1; CSIRO (2007) data (Δ) estimated using the same input data as NRC (2007). Maintenance requirements for grazing ewes (▲) calculated from CSIRO (2007) only. 143 ME requirement (MJ/day) 30 25 20 15 10 30 50 70 90 110 130 150 Ewe mature weight (kg) Figure 2. Comparison of the ME requirements for milk production in housed, mature ewes of mature weight 40 to 140 kg, with a single lamb. NRC (2007) estimates (■) and data for milk production and ewe live weight change were taken directly from their Table 15.1; CSIRO (2007) data (▲) were estimated using the same input data as NRC (2007). The remaining two requirement curves are: □ NRC (2007) estimates with efficiency of use of ME for milk production (k l) altered from a constant, as in NRC (2007), to a function of diet M/D as in CSIRO (2007); (o) NRC (2007) estimates with this adjustment and also with maintenance component increased to include a function of the level of production, as in CSIRO (2007). 144 ME requirement (MJ/day; adjusted NRC 2007) 30 25 20 15 10 10 15 20 25 30 ME requirement (MJ/day; CSIRO 2007) Figure 3. Comparison of the ME requirement of a housed, lactating ewe with a single lamb, estimated as in Figure 2 from CSIRO (2007) or from NRC (2007) with adjustment for an effect of dietary M/D on k l and for an effect of level of production on maintenance ME requirement. The linear relationship through these data points (■) does not differ from the line of equality (solid line). 145 Truly absorbed protein (g/kg DM, NRC system) 100 80 60 40 20 0 0 20 40 60 80 100 Truly absorbed protein (g/kg DM, Dutch system) Figure 4. Comparison of the truly digested and absorbed protein supply from the small intestine (g/kg DM) of dairy cows, predicted from the NRC (2001) system and the Dutch DVE/OEB system (Tamminga et al., 1994) for different pasture species and stages of growth. Constructed from the tabulated data of Yu et al. (2003): ■ Pioneer alfalfa; □ Beaver alfalfa; ▲ Climax timothy; and Δ Joliette timothy. Within a pasture species and cultivar, each symbol represents a stage of growth. The solid line is y = x. The broken line is the fitted regression in which the slope does not differ from 1, but the intercept (12.9 ± 3.61 g/kg DM) differs significantly from zero (P < 0.01). 146 Degraded protein balance (g/kg DM, NRC system) 80 60 40 20 0 -60 -40 -20 0 20 40 60 80 -20 -40 -60 Degraded protein balance (g/kg DM, Dutch system) Figure 5. Comparison of the degraded protein balance across the rumen (g/kg DM) of dairy cows, predicted from the NRC (2001) system and the Dutch DVE/OEB system (Tamminga et al., 1994) for different pasture species and stages of growth. Constructed from the tabulated data of Yu et al. (2003): ■ Pioneer alfalfa; □ Beaver alfalfa; ▲ Climax timothy; and Δ Joliette timothy. Within a pasture species and cultivar, each symbol represents a stage of growth. The solid line is the fitted regression y = 1.04x + 1.48 (R2 = 0.993; P < 0.001), which does not differ from y = x. 147 Observed live weight gain (g/day) 120 80 40 0 -80 -40 0 40 80 120 -40 -80 Predicted live weight gain (g/day) Figure 6. Comparison of the observed live weight gain of lambs grazing poor-quality summer pasture (with or without supplements of 2:1 oat grain:sunflower meal) with the gains predicted using the GrazFeed decision-support package (Freer et al., 1997). Data are re-calculated from Freer et al. (1985; □) and Freer et al. (1988; ■). The broken line is the fitted regression (Observed = 0.908*Predicted + 7.626; R2 =0.918, P < 0.001; residual standard deviation = 19.1), which does not differ from the line of equality (solid line). 148 Observed live weight gain (kg/day) 1.5 1.2 0.9 0.6 0.6 0.9 1.2 1.5 Predicted live weight gain (kg/day) Figure 7. Comparison of the observed live weight gains of crossbred steers grazing mixed grass:alfalfa pastures in western Canada, with those predicted using the GrassGro decision-support package, which incorporated the SCA (1990) nutrient requirements plus an intake predictor. The broken line is fitted regression (Observed = 1.054*Predicted – 0.068; R2 = 0.992, P < 0.001, residual standard deviation = 0.025), which does not differ from the line of equality (solid line). Recalculated from the data in Cohen et al. (2003). 149 APPLICATION OF NUTRIENT REQUIREMENT SCHEMES TO GRAZING ANIMALS H. Dove, S. R. McLennan, and D. P. Poppi Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 STRATEGIC SUPPLEMENTATION TO CORRECT FOR NUTRIENT IMBALANCES G. P. Lardy1ŧ and R. L. Endecott* ŧ * Department of Animal Sciences, North Dakota State University, Fargo 58108; and Department of Animal and Range Sciences, Montana State University, Miles City 59301 presented in Table 1. The NRC (1996) includes an interactive computer model, which allows the user to model cow size as a continuous variable and estimate nutrient requirements for an unlimited number of cow sizes. Larger cows require more total nutrients. In addition, cows with greater genetic potential for milk production also require more nutrients. Nutrient requirements for both cows were lowest during the middle trimester of gestation, following weaning (cessation of lactation). Requirements for protein and energy gradually increased throughout gestation, as a function of fetal growth. Approximately 70% of fetal growth occurs during the last third of pregnancy (NRC, 1996). This results in an increased nutrient requirement (esp., energy and protein) during the last trimester of pregnancy to support the rapidity of fetal growth during this time period. Lactation requires increased nutrient intake on the part of the lactating female. Protein and energy requirements during lactation are a function of level of milk production. Milk production typically peaks approximately 6 to 8 weeks following parturition and then declines gradually for the remainder of the lactation period. Genetic potential for milk production varies with breed type and varies widely within breeds as well, depending on selection pressure for milk production. Subsequently, levels of milk production vary accordingly. Large quantities of high-quality forage are required to maintain high levels of milk production. In situations where milk production is relatively high but forage nutrients are relatively low, the animal will typically lose weight and may fail to conceive in a timely manner. Table 2 shows the nutrient requirements for pregnant beef heifers. It is important to remember that these heifers have requirements for maintenance, growth, and pregnancy. In many cases, the requirement for growth is often overlooked, to the detriment of the nutritional status of the heifer. Nutrient requirements of growing animals, whether they are grazing or fed in confinement situations, are defined by physiological potential of the animal for growth and nutrients supplied by the forage. Of particular interest from the standpoint of animal nutrition and livestock performance is forage digestibility and protein content because nutrient supply generally limits productivity. Obviously, increases in digestibility and protein content will support greater levels of growth. For growing steers, energy and protein requirements increase with increased average daily weight gain (Table 3). INTRODUCTION Cow-calf producers in many areas of the world have access to dormant grasses during fall and winter months when lack of snow or ice cover permits grazing. Stockpiled forages (native range) represent a potentially low-cost forage resource for these producers because it requires no haying or feeding operation to deliver it to cattle (Adams et al., 1996). Cow-calf production systems that rely heavily on purchased or harvested feed inputs are also heavily reliant on fossil fuel inputs. As a consequence, these systems may be less sustainable in the long run (Heitschmidt et al., 1996). In most cases, accurate strategic supplementation is necessary to make any extensive dormant grazing program successful. Rumen-degraded protein (RDP), which is protein available to the rumen microbes, is the first limiting nutrient in dormant native range (Hollingsworth-Jenkins, 1996). Strategic supplementation will improve forage utilization and performance of cattle consuming dormant grasses. This manuscript will focus largely on strategic protein supplementation of grazing and forage-fed cattle. NUTRIENT REQUIREMENTS OF CATTLE Many texts have devoted considerable attention to describing nutrition of cattle. This paprer will simply summarize basic requirements. Readers are referred to NRC (1996) for a much more in depth review into nutrient requirements. The NRC (1996) Beef Cattle Requirements Model represents a significant change in the expression of protein requirements. The model acknowledges the protein requirements of ruminal microorganisms are different from host animal requirements. Rumen-degraded protein is the fraction of the total protein that is the primary source of N for the ruminal microorganisms. Metabolizable protein (MP) is the sum of the digestible bacterial protein and rumen-undegraded protein (RUP) present in a feedstuff. Nutrient requirements of cattle are impacted by many different factors, but the primary drivers to consider in commercial ranching operations, include cow size, stage of production (e.g. maintenance, gestation, lactation), weather (heat or cold stress), and terrain conditions (e.g. level terrain vs. hilly or mountainous terrain). Cow size and stage of production influence all types of nutrient requirements whereas terrain and weather mainly affect energy requirements. For the purposes of simplicity, only 2 cow sizes are 1 Corresponding author: [email protected] 152 between supplements were similar if cows were provided prairie hay in addition to grazing native range. No advantage was noted with feeding either biuret in a dry supplement or urea in a liquid supplement compared with feeding urea in a dry supplement. These studies were conducted with gestating cows which calved during the trials, which makes interpretation difficult, because protein requirements change dramatically from gestation to lactation. Urea and biuret supply degradable protein needed by the ruminal microorganisms, but do not supply metabolizable protein required by the animal. The use of 15% urea:85% corn, 100% soybean meal, 10% corn:40% soybean meal:50% urea, or 14% corn:36% blood meal:50% urea as supplements for low-quality native forage was investigated by Petersen et al. (1985). They found OM digestibility, bacterial N flow to the abomasum, feed N flow to the abomasum, and microbial efficiency were not different among these supplements. In a companion N balance study (Peterson et al., 1985), no supplement, soybean meal, soybean meal and urea, and urea and blood meal were fed with low-quality native forage. Supplementation increased digestibility of DM and NDF, and N retention was greater for steers fed soybean meal or blood meal-urea. They concluded that a blend of blood meal and urea was an effective supplement for cattle consuming low quality native grass. Köster et al. (2002) conducted four trials to evaluate the effect of increasing level of urea in protein supplements on performance of prepartum cows grazing dormant native range or cannulated steers fed low-quality dormant native range. Forage intake and digestibility were not affected by urea level when supplements were fed to cannulated steers. At the higher levels of urea inclusion, cows lost more weight and body condition prepartum than those fed no or lower levels of urea. Pregnancy rate was generally unaffected by supplemental urea level; however, there was a tendency for it to be lower in one study at the 40% (of supplemental RDP) level. They concluded that urea could replace from 20 to 40% of the supplemental RDP in 30% CP supplements without significantly altering either supplement palatability or cow performance. It should be noted that some supplement refusal was noted when cows were fed supplements containing 60% of the supplemental DIP from urea. These problems may be overcome with additional research, but should be a practical consideration when formulating supplements. Farmer et al. (2004) investigated the interaction between urea level and supplementation frequency. They noted increased supplement refusal when greater levels of urea were included in the supplement and when supplements were fed on alternate days. Additionally, increasing level of urea in the supplement resulted in increased BW loss in cows grazing dormant tallgrass prairie. Those researchers also noted a tendency for lower pregnancy rates at higher inclusion rates of urea in one of the three prepartum cow trials. Supplements used in Köster et al. (2002) and Farmer et al. (2004) were provided at a rate of 2.16 and 1.64 kg/d of DM, respectively. It is unclear whether similar results would be observed if supplements were delivered in a smaller package size with less carrier. PROTEIN SUPPLEMENTATION Non-Protein Nitrogen Supplements If a source of ammonia is all that is necessary to improve digestion and intake of low-quality forages, nonprotein nitrogen (NPN) should provide adequate ammonia levels for this purpose. Clanton (1978) summarized several experiments conducted on native sandhills range using NPN in protein supplements and found that NPN was not as effective as all natural protein supplements as a source of protein for growing calves wintered on native range. Possibly the MP requirement of these calves was great enough that they responded to the additional escape protein. Calves supplemented with biuret had similar rates of gain compared with calves supplemented with urea (Clanton, 1978), indicating that biuret provided no additional benefit over and above urea. Currier et al. (2004) compared daily and alternate-day urea and biuret supplementation of low-quality (4% CP) grass straw to no supplementation for wethers and prepartum cows. Performance was improved when animals received supplement, but neither source of NPN nor frequency of supplementation of NPN impacted digestibility, N balance, or N retention by wethers, or weight and body condition change of prepartum cows. The study did not include a positive control (natural protein supplement), which leaves one to speculate whether performance would have been further improved if a natural protein source were included for comparison. When urea replaced soybean meal in supplements offered to cannulated steers fed low-quality range forage, digestion of DM and OM decreased, but synthesis of microbial protein was relatively unchanged (Kropp et al., 1977). In this study, soybean meal was replaced isocalorically and isonitrogenously with urea and ground sorghum. It is possible that the sorghum in the supplements influenced DM and OM digestibility due to negative associative effects of starch digestion on fiber utilization (Mertens and Loften, 1980; Chase and Hibberd, 1987). Ground sorghum levels ranged from 9.5 to 19.6% of the diet and DMI averaged 4.6 kg/d. Starch content of the sorghum and starch intake was not reported. Mlay et al. (2003) supplemented dairy heifers consuming low-quality hay with either urea or soybean cake at two levels. At each level, the urea and soybean cake were isonitrogenous, and urea was added via rumen cannulae. Nitrogen supplementation, regardless of source, linearly increased forage intake and digestibility, which led those authors to conclude that soybean cake and urea were equivalent sources of supplemental N. Soybean cake was suggested to be slightly superior to urea due to its slower degradation in the rumen, which probably better matched the degradation rate of the low-quality hay. The effectiveness of biuret or urea in dry or liquid supplements compared with all natural protein supplements for cows grazing dormant winter range was investigated by Rush and Totusek (1976). They found that cows fed the all natural protein supplement had less weight loss than isonitrogenous blends of molasses and urea when cows were foraging on native range only. However, weight losses 153 Other workers have not reported increased intake when protein supplements were offered to cattle consuming lowquality forages (Krysl et al., 1989; Hollingsworth-Jenkins et al., 1996; Reed et al., 2004). In studies that report no effect on forage intake in response to protein supplementation, forage CP levels are generally greater than in studies that do report a response. This should be intuitive, however, it is difficult to generalize when responses may or may not be observed, simply based on forage CP level. Basal forage intake likely contributes to some of the variation in the responses observed in these studies. For example, in the case of Hollingsworth-Jenkins (1996), intake by control cows was approaching 2.0% of BW, while intake of control animals in studies that have reported responses are sometimes considerably lower, such as in the work of DelCurto (1990b) in which intake averaged 0.49% of BW and DelCurto et al. (1990a) in which intake averaged 0.87% of BW. Clearly, more research is needed to provide ranchers with the information necessary to make more strategic decisions regarding supplementation. Responses in digestibility to protein supplements are also variable. Some workers have reported increased digestibility when protein supplements were fed (Fleck et al., 1988; Sunvold et al., 1991; Horney et al., 1996), whereas others (Kartchner, 1980; Petersen et al., 1985; Köster et al. 1996) reported no differences in digestibility with supplementation. Villalobos et al (1997b) reported increases in digestibility with protein supplementation when measured in metabolism studies, but differences in digestibility were not detected with grazing cows fed similar supplements (Villalobos et al., 1997a). Some of these differences are attributable to methods used to determine digestibility in the various studies (Galyean et al., 1986; Cochran et al., 1987). Unfortunately, there is a great deal of variation in response to feeding supplemental protein in grazing or forage feeding situations. Factors which affect this variation include environmental influences such as temperature and snow cover, forage quality, and physiological status of the animal. Additional research which focuses on understanding mechanisms which elicit responses in intake and digestibility are needed in order to allow for improved decision making regarding protein supplementation. Provision of microbial growth factors. The effect mediated by natural protein supplements may be due, in part, to the provision of branched-chain VFA through the fermentation of leucine, isoleucine, and valine. These VFA are either required or highly stimulatory to cellulolytic bacteria (Mackie and White, 1990). No differences in intake, digestibility, or microbial CP production were observed in steers given supplemental branched-chain VFA, indicating that the supply of branched-chain VFA were not limiting or had no effect on microbial fermentation with the diets tested (McCollum et al., 1987). Natural protein supplements may also supply limiting AA. Blends of urea and DL-methionine, urea and ammonium sulfate, or soybean meal as supplements for ruminally cannulated crossbred beef cows fed a 75:25 blend of grass hay and barley straw were investigated by Clark and Petersen (1988). In situ rate of fermentation was increased with the methionine treatment. Similar treatments Urea and other forms of NPN may be used to provide a portion of the supplemental RDP for cattle grazing dormant native range or offered other low-protein forages. However, they do not seem to be adequate as the sole source of supplemental RDP for cattle consuming low-quality forage diets. The stimulatory effect of natural protein, peptides, and (or) AA in vitro indicate that rumen microbes require natural protein and (or) AA for optimum growth. The results of supplementation trials comparing NPN to natural protein supplements indicate that urea alone is not an effective N supplement for cattle grazing dormant native range unless a portion of the N is supplied by protein or peptides. Effects of Feeding Natural Protein Supplements Several broad categories of effects have been reported when supplementing low-quality forages with natural protein-based supplements. These include increases in forage intake, increased digestibility, and increased rate of passage. Variation in responses is common and not all effects are observed in each trial. Most responses to protein supplements were observed when forage CP levels were below 7% CP (Paterson et al., 1994). In most studies, protein supplements were based on oilseed meals, which contain both RDP and RUP; thus, ascertaining the biological mechanism responsible for eliciting responses difficult. Additionally, protein deficiency has not always been demonstrated. Because of differential responses between urea-based supplements and natural protein, a factor other than rumen ammonia provided by natural protein supplements may be responsible for the effects observed when natural protein supplements are fed. These factors could include: 1) provision of N in a form other than ammonia (i.e., AA or peptides), which are possibly stimulatory toward the rumen microbial population; 2) provision of some other growth factor required by the ruminal bacteria, which could include branched-chain VFA; 3) provision of RUP, which helps to meet the MP requirements of the animal; or 4) stimulation of ruminal kinetics, allowing increased flow of nonnitrogenous and N-containing compounds (Petersen, 1987). A review of literature pertaining to supplementation programs for ruminants grazing dormant native range indicates that responses to supplemental protein have been highly variable and vary from year to year. This fact alone argues for more complete data regarding seasonal effects on nutrient content, protein degrability, and microbial efficiency. Factors contributing to or influencing this variation include weather conditions, which includes temperature and snow cover (Kartchner, 1980), animal factors such as cow age (Adams et al., 1986), and differences in forage quality from year to year (Soder et al., 1995; Kartchner, 1980). Protein supplementation of forages low in CP increased intake (Hennessy et al., 1983; McCollum and Galyean, 1985; Gilbery et al., 2006). Mathis et al (1999) fed increasing levels of soybean meal to beef steers consuming 5.3% CP native grass hay. Forage OM intake increased until supplemental soybean meal intake reached 0.16% of BW. 154 detected in DM intake, possibly because the basal diet was 8.75% CP. Supplements varying in ruminal degradability (25 or 50% RUP as a % of CP) were fed to 2 groups of lactating spring calving cows (Dhuyvetter et al., 1992). Cows were split into groups based on calving date. Early calving cows grazed native range and were fed 5.4 kg of medium-quality grass hay (10.5% CP); late calving cows grazed native range only. Late calving cows had similar weight losses, regardless of which supplement was fed. Early calving cows fed the 50% RUP supplement lost 39 kg less weight than cows fed the 25% RUP supplement. Dhuyvetter et al. (1992) concluded that the late calving cows did not respond to RUP in the same manner as the early calving cows because RDP may have been limiting. Response to RUP was dependent on inclusion level in the diet with lactating ewes (Loerch et al., 1985). When compared to a diet supplemented with soybean meal, blood meal fed at 3.3% of diet DM increased milk production from 2.5 to 3.2 kg/d. In a second study, when the blood meal was included at 6.8% of the diet, feed intake, and milk production tended to be lower as compared with the diet supplemented with soybean meal. The authors attributed these responses to the poor palatability of the diet containing high levels of blood meal. Dry matter intake was 3.0 kg when soybean meal was used as the supplement and 2.7 kg when 6.8% blood meal was included in the diet. Milk production by Hereford cows increased with increasing level [0 to 21 g supplement/kg body weight 0.75] of protein supplement was offered with fed a low-quality tropical grass hay (2.7% CP; Lee et al., 1985). Lee et al. (1985) used a protein supplement made up of a blend of cottonseed meal, fishmeal, meat meal, and mineral mix at 75.5%, 9.45%, 9.45%, and 5.6% of the supplement DM, respectively. The authors stated that 53% of the N in the supplement remained after a 15-h in situ incubation. Nonsupplemented cows lost 2.56 kg/d whereas cows fed the highest level of supplementation gained 0.15 kg/d. Growth rates of calves that nursed cows receiving higher levels of supplement were also higher. Sletmoen-Olson et al. (2000) fed lactating beef cows increasing levels of RUP. Control cows (nonsupplemented) had greater forage intake than supplemented cows. No differences between supplemented cows were noted. No differences in total intake (forage plus supplement) were noted. Supplemented cows had greater BW during lactation than non-supplemented controls, but no differences existed between supplemented groups. Encinias et al. (2005) offered no supplement, an energy control, RDP supplement, or a supplement containing both RDP and RUP to lactating beef cows consuming moderatequality forage (9.6% CP). No differences in cow performance, milk production, or calf weight gain were noted in the study, indicating that MP supply from the forage and microbial protein was adequate. Waterman et al. (2006) supplemented young postpartum beef cows grazing dormant native range with one of three 36% CP supplements: 1) RDP (cottonseed meal-based), 2) 50:50 RDP:RUP supplement (cottonseed meal and feather meal), and 3) 50:50 RDP:RUP supplement plus 100 g/d propionate salt. Cows supplemented with RUP had were imposed on weaned heifer calves, but no differences were found between the soybean meal and urea-methionine supplements. The use of a blend of sugar beet pulp and DLmethionine or soybean meal as a supplement for gestating beef cows grazing native range was investigated by Lodman et al. (1990). The methionine treatment did not support the same level of performance as the soybean meal treatment. Momont et al. (1993), in a similar trial, compared soybean meal to a blend of urea, corn, and methionine for cows grazing dormant winter range. They also found that the urea, corn, and methionine treatment did not support the same level of cow performance as the soybean meal treatment. Feeding to meet the metabolizable protein needs. Patterson et al. (2003) evaluated the effects of feeding supplements balanced to meet the CP needs or the MP requirements of primiparous beef heifers. The research involved 2,120 primiparous heifers maintained in a commercial ranching environment in the Nebraska Sandhills. Heifers supplemented to meet their MP requirements had similar BW and BCS at the end of the feeding periods. However, heifers fed to meet their MP requirements were heavier at the time of pregnancy diagnosis as two-year olds. Pregnancy rates were also greater (91 vs. 86%) for heifers fed to meet their MP requirements compared with heifers fed to meet their CP needs. Supplementation of Lactating Ruminants Lactating cows grazing smooth bromegrass pastures and fed increasing levels of supplemental RUP had increased milk production and increased calf weight gain (Blasi et al., 1991). No response was noted when cows grazed big bluestem, however. In situ analysis of forage samples suggested that RUP content of smooth bromegrass was less than that of big bluestem. The effect of time of initiation of protein supplementation on spring-calving cow performance was examined by Sowell et al. (1992). Protein supplementation began either pre-calving or post-calving. Cows that were supplemented pre-calving had less spring weight loss and greater prebreeding BCS than cows that did not receive supplement until after calving. However, no differences were noted in reproductive performance. Ovenell et al. (1991) reported no interactions between lactational status and supplement type (soybean meal, wheat middlings, or corn-soybean meal) for hay intake, digestibility, or particulate passage rate. Differences in intake were detected, 2.1% and 1.9% of BW, respectively between lactating and pregnant cows. No differences were observed in DM digestibility or particulate passage rate. Hunter and Magner (1988) supplemented lactating primiparous cows with formaldehyde-treated casein. They found no differences in milk production during the first 8 weeks of lactation; however, during the second half of lactation, supplemented heifers produced less milk than unsupplemented heifers. Longer periods of anestrus were noted for unsupplemented heifers. No differences were 155 change, or cow BCS change, most often the economics of the situation justify alternate day or twice weekly feeding of protein supplements. increased milk production yet returned to estrus sooner after calving compared with cows that received the RDP supplement. Overall pregnancy rates were similar for all cows. Cows fed the propionate salt-containing supplement also produced more milk and returned to estrus quicker than cows fed RDP. Cows fed the propionate salt also had the shortest glucose half-life after a glucose tolerance test, suggesting that nutrient partitioning can be influenced by supplement composition. The most consistent response from supplementing natural protein to the lactating beef cow was increased milk production. However, this is not observed in all studies and may be related to supply of essential AA in the supplemental source of protein. In many cases, calf gain increased due to increased milk production when lactating cows were supplemented with natural protein. Interaction Between Supplemental Degradable Protein and Energy Level of supplemental protein fed when feeding cereal grain supplements also seems to interact with forage quality. Bodine et al. (2000) fed supplemental corn at 0 or 0.75% of BW to beef steers consuming low-quality (6% CP) grass hay. When supplemental RDP was provided, the negative associative effects of corn feeding were less apparent. Additional research at Oklahoma State University investigated the relationship between supplemental RDP and energy supplementation for cattle grazing dormant forages (Bodine and Purvis, 2003). Those researchers noted improved performance in growing steers grazing dormant tallgrass prairie when the supply of RDP was balanced with total dietary TDN supply. Baumann et al. (2004) reported interactions between source of energy supplement (corn vs. soyhulls) and provision of a RDP supplement when cattle were fed a low-quality forage diet. Supplementation Frequency Protein supplements do not need to be provided on a daily basis. Many reports in the literature indicate only small changes in forage utilization, cow weight change, or cow BCS change when protein supplements are provided as often as daily or as infrequently as weekly (Kartchner and Adams, 1982; Beaty et al., 1994; Farmer et al., 2001). Farmer et al. (2001) supplemented 2, 3, 5, or 7 times per week and found that forage use was improved by supplementing more frequently, but that impacts on animal performance were likely minimal. Beaty et al. (1994) concluded that while digestion characteristics were favored slightly by daily supplementation, supplementing 3 times per week decreased labor with only minimal consequences on cow performance. Huston et al. (1999) supplemented cows 0, 1, 3, or 7 days per week and found that supplemented cows lost less BW and body condition than non-supplemented cows, and less variability in supplement intake was observed in cows fed once or 3 times per week compared with those fed daily. Those authors attributed this to the aggressive competition during the short consumption period for the daily-supplemented animals. Decreased variation in forage intake and live BW change were also noted in the cows fed less frequently. Huston et al. (1999) concluded that protein entering the cow’s system in large amounts does not require immediate use to have lingering value. Krehbiel et al. (1998) investigated the effects of soybean meal supplementation frequency on portal and hepatic flux of nutrients in ewes fed 7.5% CP bromegrass hay. Soybean meal was supplemented daily or every third day to provide 80 g CP daily and a non-supplemented control was included in the treatment design. They noted that net flux of nutrients was generally unchanged by frequency of supplementation. However, urea removal by the portal drained viscera was greater when supplement was provided every third day, which may have implications for the N efficiencies of the ruminal microorganisms. Schauer et al. (2005) investigated the effect of supplementation frequency (daily or every six days) on performance, grazing behavior, and variation in supplement intake. No differences in performance, grazing behavior, or variation in supplement intake were noted. Even when small differences existed in forage utilization, cow weight FAT SUPPLEMENTATION OF GRAZING RUMINANTS Fat supplementation of ruminants has been investigated as a means to increase diet energy density, to modify and improve reproductive function, or to alter fatty acid composition of end products. Results of fat supplementation have been variable. Two comprehensive reviews of fat supplementation that readers are encouraged to reference are Funston (2004) and Hess et al. (2008). Feeding fat can increase diet energy density, and limiting supplemental fat to less than 2% of diet DM will decrease the potential negative associative effects for cows consuming forages (Hess et al., 2008). Positive reproductive responses to fat supplementation are probably due to changes in fatty acid status rather than diet energy density (Funston, 2004, Hess et al., 2008). Supplementing replacement heifers for 60 to 90 days before breeding has positive effects, but supplementing fat postpartum did not Fat impact pregnancy rate (Hess et al., 2008). supplementation of the gestating dam can improve fatty acid status of the neonate, which can improve survivability in harsh environments (Hess et al., 2008). STRATEGIC SUPPLEMENTATION TO MANIPULATE GRAZING DISTRIBUTION In many areas of the western U.S., grazing distribution has become an increasingly important concern, especially in environmentally sensitive areas (e.g., riparian areas). The use of supplementation as a means to manipulate livestock movements has been investigated in a series of studies conducted by scientists in Montana and Oregon. Bailey et al. (2008) compared the use of salt or low-moisture molasses blocks as a means to improve pasture utilization in areas that were historically underutilized. Low-moisture molasses blocks were more useful in implementing this 156 conditions, such as amount and timing of precipitation, date of frost at the beginning and end of the growing season, and other conditions that affect plant growth. Table 4 gives the seasonal changes in nutritive value of native range in southwestern North Dakota. Crude protein and digestibility of native range forage decreased as the growing season advances. In some production systems, this may make supplementation necessary for a portion of the grazing season. Table 5 illustrates the seasonal changes in nutritive value of subirrigated meadow in the Nebraska Sandhills. Table 6 gives the seasonal changes in nutritive value of native range in the Nebraska Sandhills. The nutrient analysis from these forage samples clearly underly the importance of understanding the type of forage and nutrient characteristics of each plant mix throughout the grazing season. The subirrigated meadow is predominantly cool season species whereas the upland range is predominantly warm season. Dominant grass species on the native upland range sites were as follows: little bluestem (Schizachyrium scoparium [Michx.] Nash), prairie sandreed (Calamovilfa longifolia [Hook.] Scribn.), sand bluestem (Andropogon gerardii var. paucipilus [Nash] Fern.), switchgrass (Panicum virgatum L.), sand lovegrass (Eragrostis trichodes [Nutt.] Wood), indiangrass (Sorghastrum nutans [L.] Nash), and blue grama (Bouteloua gracilis [Willd. Ex H.B.K.] Lag. Ex Griffiths). Common forbs and shrubs include western ragweed (Ambrosia psilostachya DC.) and leadplant (Amorpha canescens [Nutt.] Pursh). Dominant vegetation on the subirrigated meadows consisted of smooth bromegrass (Bromus inermis Leyss), redtop (Agrostis gigantea Roth.), timothy (Phleum pratense L.), slender wheatgrass (Elymus trachycaulus [Link] Gould ex Shinners), quackgrass (Elytrigia repens [L.] Nevski.), Kentucky bluegrass (Poa pratensis L.), prairie cordgrass (Spartina pectinata Link), and several species of sedges (Carex) and rushes (Juncus). Less-abundant grass species were big bluestem (Andropogon gerardii var. gerardii Vitman), indiangrass (Sorghastrum nutans [L.] Nash), and switchgrass (Panicum virgatum L.). Abundant legumes included red clover (Trifolium pretense L.). Differential responses to protein supplementation with cool and warm season forages have also been noted. Work by Bohnert et al. (2007) with bluegrass straw (C3; 5.7% CP) and tall-grass prairie hay (C4; 6.3% CP) indicated larger increases in intake with the C4 (47% increase) compared with the C3 (7% increase) when protein supplement was provided. Additionally, DM digestibility responded similarly with digestibility increasing 12% with C4 and 9% with C3 forage. These results indicate the need for additional research to investigate the plant and animal mechanisms involved in the response. Additional data regarding seasonal changes in RDP content of native forage under various growing conditions is necessary to better understand the need for supplementation (Lardy et al., 2004). Cost-effective decisions regarding protein supplementation of grazing livestock ultimately requires knowledge of both the animal’s nutrient requirements as well as seasonal changes in forage nutrients. strategy than was salt. When low-moisture molasses blocks were used, cows grazed higher elevations and further horizontally from water than when salt was used. Bailey and Jensen (2008) compared the use of range cake (cubes) and low-moisture molasses blocks as a means to improve grazing distribution in Montana foothills grazing areas. Cows fed the low-moisture blocks were observed to graze steeper areas than the cows fed range cake. Those scientists also reported differences in forage utilization when the 2 supplement types were compared. Forage utilization decreased as slope increased to a greater degree when range cake was fed than when low-moisture molasses block was fed. Cows spent more time within 100 m of sites with a low-moisture molasses block than sites where range cubes were offered. Ganskopp (2001) investigated the use of manipulation of sites where cattle accessed water and salt as a means to improve grazing distribution. This work moved water sources and salt locations either together or separately to investigate which was more effective in achieving improved pasture utilization. Results indicated cattle were more often found near water sites than sites where salt was offered. Centers of activity moved further when water locations were changed than when salt locations were changed. When salt locations were changed, cattle used the new sites for approximately 2 d and then began to move back to more familiar parts of the pasture. SUPPLEMENT TYPE CONSIDERATIONS Bowman and Sowell (1997) reviewed the effects of supplement type and noted the following observations. The percentage of non-eaters (animals not consuming supplement) was similar for blocks and dry supplements (approximately 15%), but slightly greater for liquid supplements (23%). In studies that directly compared hand-fed supplements and self-fed supplements, the percentage of non-eaters was greater for the self-fed supplements (19%) than the hand-fed supplements (5%). The coefficient of variation for individual supplement consumption averaged 79% for blocks, 41% for dry, and 60% for liquid supplements. In some formulation processes, chemical hardeners are used as a means to control intake of block supplements. As one would expect, as block hardness increases block intake decreases. However, the coefficient of variation in supplement intake increases as block hardness increases (Zhu et al., 1991). Livestock producers should take supplement type into consideration as they develop strategic supplementation programs, realizing that each has inherent advantages and disadvantages. What works well for one ranch may not work well for another due to a variety of factors, including animal behavior, ranch resources, management ability, and resource utilization goals. SEASONAL CHANGES IN NUTRIENT QUALITY OF NATIVE RANGE Numerous research stations have investigated seasonal changes in nutrient quality of native range over the growing season. Year-to-year variation in forage quality occurs. Much of this variation is related to environmental 157 performance. Protein supplements may be used as a strategic means of enhancing forage utilization in areas with rough terrain. This is a particularly useful tool in environmentally sensitive rangeland areas. The role of supplementation in developmental programming is just beginning to be explored. Early results indicate potential to manipulate carcass quality and progeny reproductive success with strategic supplementation. Strategic supplementation will improve beef cattle performance and allow utilization of dormant native range with minimal cost. DEVELOPMENTAL PROGRAMMING A review on strategic supplementation would not be complete without some discussion of developmental programming. Developmental programming is a term coined to describe the response that a fetus has to an inutero insult or stress (Barker, 2004; Wu et al., 2006; Reynolds et al., 2010). In relation to this manuscript, it is also used to describe the potential effects of strategic supplementation during gestation (or lack thereof) and subsequent effects on offspring growth, development, and productivity. In some cases, these effects are observed well after supplementation ends (Funston et al., 2010). Perhaps the best evidence for this phenomenon in beef cattle is evidenced by the work of Martin et al. (2007). In this work, a 2 x 2 factorial was used to investigate the effects of late gestation protein supplementation and early lactation nutrition strategy. No effects were noted due to early lactation nutrition. However, dams supplemented with protein during the last one-third of pregnancy gave birth to heifer calves that subsequently had increased pregnancy rate and increased percentages calving in the first 21 d of the calving season compared with heifers from nonsupplemented dams. Additional work from the same laboratory indicates that fewer heifers from nonsupplemented dams attained puberty before the first breeding season compared with heifers from supplemented cows in a subsequent study (Funston et al., 2008). Supplementation also increased the percentage of carcasses grading USDA Choice when those progeny were finished. Additional research is necessary to determine the key times during gestation when protein supplementation may allow these responses to manifest themselves and to fully understand the mechanisms behind them. This work should involve both an understanding of the underlying biology at work, but also encompass a systems approach that allows producers insight into appropriate, strategic, and management decisions regarding supplementation to improve profitability. LITERATURE CITED Adams, D. C., T. C. Nelsen, W. L. Reynolds, and B. W. Knapp. 1986. Winter grazing activity and forage intake of range cows in the northern Great Plains. J. Anim. Sci. 62:1240-1246. Adams, D. C., R. T. Clark, T. J. Klopfenstein, and J. D. Volesky. 1996. Matching the cow with forage resources. Rangelands. 18: 57-62 Bailey, D. W. and D. Jensen. 2008. Method of supplementation may affect cattle grazing patterns. Rangeland Ecol. Manage. 61:131–135. Bailey, D. W, H. C. VanWagoner, R. Weinmeister, and D. Jensen. 2008. Comparison of low-moisture blocks and salt for manipulating grazing patterns of beef cows. J. Anim. Sci.86:1271-1277. Barker, D. J. 2004. Developmental origins of well being. Philos. Trans. Royal Soc. (Lond.) 359:1359-1366. Baumann, T. A, G. P. Lardy, J. S. Caton, and V. L. Anderson. 2004. Effect of energy source and ruminally degradable protein addition on performance of lactating beef cows and digestion characteristics of steers. J. Anim. Sci. 82:2667-2678. Beaty, J. L., R. C. Cochran, B. A. Lintzenich, E. S. Vanzant, J. L. Morrill, R. T. Brandt, Jr., and D. E. Johnson. 1994. Effect of frequency of supplementation and protein concentration in supplements on performance and digestion characteristics of beef cattle consuming low quality forages. J. Anim. Sci. 72:24752486. Blasi, D. A., J. K. Ward, T. J. Klopfenstein, and R. A. Britton. 1991. Escape protein for beef cows: III. Performance of lactating beef cows grazing smooth brome or big bluestem. J. Anim. Sci. 69:2294-2302. Bodine, T. N. and H. T. Purvis, II. 2003. Effects of supplemental energy and/or degradable intake protein on performance, grazing behavior, intake, digestibility, and fecal and blood indices by beef steers grazed on dormant native tallgrass prairie. J. Anim. Sci. 81:304317. Bodine, T. N., H. T. Purvis, II, C. J. Ackerman, and C. L. Goad. 2000. Effects of supplementing prairie hay with corn and soybean meal on intake, digestion, and ruminal measurements by beef steers. J. Anim. Sci. 78:31443154. CONCLUSIONS A wide variety of factors influence the success of supplementation programs. Knowledge of and familiarity with nutrient content of grazed forages is important for designing successful strategic supplementation programs. Without some knowledge of forage quality, it is almost impossible to make informed supplementation decisions. Rumen-degraded protein is limiting in low-quality, dormant forages and native range. Use of supplements high in RDP should be considered in these situations. Urea or other NPN sources generally do not produce satisfactory results when used as the sole source of RDP. They may be used in combination with natural protein sources with a greater degree of success due, in part, to the provision of microbial growth factors that enhance microbial fermentation of lowquality forages. Lactating beef cows may need supplemental RUP when forage quality is low and (or) when milk production is high. Research data indicate that protein supplements may be provided infrequently (e.g., once or twice weekly) with acceptable levels of livestock 158 Bohnert, D.W., T. DelCurto, A. A. Clark, M. L. Merrill, S. J. Falck, and D. L. Harmon. 2007. Protein supplementation of ruminants consuming low-quality cool- or warm-season forage: differences in intake and digestibility. Proc. West. Sec. Amer. Soc. Anim. Sci. 58:217-220. Bowman, J. G. and B. F. Sowell. 1997. Delivery method and supplement consumption by grazing ruminants: a review. J. Anim. Sci. 75:543-550. Chase, C. C. and C. A. Hibberd. 1987. Utilization of lowquality native grass hay by beef cows fed increasing levels of corn grain. J. Anim. Sci. 65:557-566. Clanton, D. C. 1978. Non-protein nitrogen in range supplements. J. Anim. Sci. 47:765-779. Clark, C. K. and M. K. Petersen. 1988. Influence of DLmethionine supplementation on growth, ruminal fermentation and dilution rate in heifers. J. Anim. Sci. 66:743-749. Cochran, R. C., E. S. Vanzant, K. A. Jacques, M. L. Galyean, D. C. Adams, and J. D. Wallace. 1987. Internal markers. Pages 39-48 In: Proc. Grazing Livst. Nutr. Conf. Univ. of Wyoming, Laramie, WY. Currier, T. A., D. W. Bohnert, S. J. Falck, and S. J. Bartle. 2004. Daily and alternate day supplementation of urea or biuret to ruminants consuming low-quality forage: I. Effects on cow performance and the efficiency of nitrogen use in wethers. J. Anim. Sci. 82:1508-1517. DelCurto, T., R. C. Cochran, D. L. Harmon, A. A. Beharka, K. A. Jacques, G. Towne, and E. S. Vanzant. 1990a. Supplementation of dormant tallgrass-prairie forage: I. Influence of varying supplemental protein and(or) energy levels on forage utilization characteristics of beef steers in confinement. J. Anim. Sci. 68:515-531. DelCurto, T., R. C. Cochran, T. G. Nagaraja, L. R. Corah, A. A. Beharka, and E. S. Vanzant. 1990b. Comparison of soybean meal/sorghum grain, alfalfa hay and dehydrated alfalfa pellets as supplemental protein sources for beef cattle consuming dormant tall grass prairie forage. J. Anim. Sci. 68:2901-2915. Dhuyvetter, D.V., M. K. Petersen, R. P. Ansotegui, R. A. Bellows, B. Nisley, R. Brownson, and M. W. Tess. 1992. Reproduction efficiency of range beef cows fed differing sources of protein during gestation and differing quantities of ruminally undegraded protein prior to breeding. Proc. West. Sec. Amer. Soc. Anim. Sci. 43:59-62. Encinias, A. M., G. P. Lardy, J. L. Leupp, H. B. Encinias, L. P. Reynolds and J. S. Caton. 2005. Efficacy of using a combination of rendered protein products as an undegradable intake protein supplement for lactating, winter-calving, beef cows fed bromegrass hay. J. Anim. Sci. 83:187-195. Farmer , C. G., R. C. Cochran, D. D. Simms, E. A. Klevesahl, T. A. Wickersham, and D. E. Johnson. 2001. The effects of several supplementation frequencies on forage use and the performance of beef cattle consuming dormant tallgrass prairie forage. J. Anim. Sci. 79:22762285. Farmer, C. G., B. C. Woods, R. C. Cochran, J. S. Heldt, C. P. Mathis, K. C. Olson, E. C. Titgemeyer, and T. A. Wickersham. 2004. Effect of supplementation frequency and supplemental urea level on dormant tallgrass-prairie hay intake and digestion by beef steers and prepartum performance of beef cows grazing dormant tallgrassprairie. J. Anim. Sci. 82:884–894. Fleck, A. T., K. S. Lusby, F. N. Owens, and F. T. McCollum. 1988. Effects of corn gluten feed on forage intake, digestibility, and ruminal parameters of cattle fed native grass hay. J. Anim. Sci. 66:750-757. Funston, R. N. 2004. Fat supplementation and reproduction in beef females. J. Anim. Sci. 82:E154-E161. Funston, R. N., J. L. Martin, D. C. Adams, and D. M. Larson. 2008. Effects of winter grazing system and supplementation during late gestation on performance of beef cows and progeny. Proc. West. Sec. Amer. Soc. Anim. Sci. 59:102-105. Funston, R. N., D. M. Larson, and K. A. Vonnahme. 2010. Impacts of maternal nutrition on conceptus growth and offspring performance: Implications for beef cattle production. J. Anim. Sci. 88:E205-E215. Galyean, M. L., L. J. Krysl, and R. E. Estell. 1986. Markerbased approaches for estimation of fecal output and digestibility in ruminants. Pages 96-113 In: Feed Intake by Beef Cattle: Symposium F.N. Owens (ed) Okla. Agric. Exp. Sta. MP-121. Stillwater, Okla. Ganskopp, D. 2001. Manipulating cattle distribution with salt and water in large arid-land pastures: A GPS/GIS assessment. Appl. Anim. Behav. Sci. 73:251–262. Gilbery, T.C., G. P. Lardy, S. A. Soto-Navarro, M. L. Bauer, and J. S. Caton. 2006. Effects of corn condensed distillers solubles supplementation on ruminal fermentation, digestion, and in situ disappearance in steers consuming low-quality hay. J. Anim. Sci. 84:1468-1480. Heitschmidt, R. K., R. E. Short, and E. E. Grings. Ecosystems, sustainability, and animal agriculture. J. Anim. Sci. 1996. 74:1395-1405. Hennessy, D. W., P. J. Williamson, J. V. Nolan, T. J. Kempton, and R. A. Leng. 1983. The roles of energy- or protein-rich supplements in the subtropics for young cattle consuming basal diets that are low in digestible energy and protein. J. Agric. Sci. (Camb.) 100:657-666. Hess, B. W., G. E. Moss, and D. C. Rule. 2008. A decade of developments in the area of fat supplementation research with beef cattle and sheep. J. Anim. Sci. 86:E188-E204. Hollingsworth-Jenkins, K. J., T. J. Klopfenstein, D. C. Adams, and J. B. Lamb. 1996. Ruminally degradable protein requirement of gestating beef cows grazing native winter sandhills range. J. Anim. Sci. 74:13431348. Horney, M. R., T. DelCurto, M. M. Stamm, R. K. Bailey, and S. D. Brandyberry. 1996. Early-vegetative tall fescue hay vs alfalfa hay as a supplement for cattle consuming low-quality roughages. J. Anim. Sci. 74:1959-1966. 159 Hunter, R. A. and T. Magner. 1988. The effect of supplements of formaldehyde-treated casein on the partitioning of nutrients between the cow and calf in lactating Bos Indicus x Bos taurus heifers fed a roughage diet. Aust. J. Agr. Res. 39:1151-1162. Huston, J. E., H. Lippke, T.D.A. Forbes, J. W. Holloway, and R. V. Machen. 1999. Effects of supplemental feeding interval on adult cows in western Texas. J. Anim. Sci. 77:3057-3067. Johnson, J. A., J. S. Caton, W. Poland, D. R. Kirby, and D. V. Dhuyvetter. 1998. Influence of season on dietary composition, intake, and digestion by beef steers grazing mixed-grass prairie in the northern Great Plains. J. Anim Sci 76:1682-1690. Kartchner, R. J. 1980. Effects of protein and energy supplementation of cows grazing native winter range forage on intake and digestibility. J. Anim. Sci. 51:432438. Kartchner , R. J. and D. C. Adams. 1982. Effects of daily and alternate day feeding of grain supplements to cows grazing fall-winter range. Mont. Agric. Exp. Sta. Bull. No. 1281:308-311. Köster, H. H., R. C. Cochran, E. C. Titgemeyer, E. S. Vanzant, I Abdelgadir, and G. St-Jean. 1996. Effect of increasing degradable intake protein on intake and digestion of low-quality, tallgrass-prairie forage by beef cows. J. Anim. Sci. 74:2473-2481. Köster, H. H., B. C. Woods, R. C. Cochran, E. S. Vanzant, E. C. Titgemeyer, D. M. Grieger, K. C. Olson, and G. Stokka. 2002. Effect of increasing proportion of supplemental N from urea in prepartum supplements on range beef cow performance and on forage intake and digestibility by steers fed low-quality forage. J. Anim. Sci. 80:1652-1682. Krehbiel, C. R., C. L. Ferrell, and H. C. Freetly. 1998. Effects of frequency of supplementation on dry matter intake and net portal and hepatic flux of nutrients in mature ewes that consume low-quality forage. J. Anim. Sci. 76:2464-2473. Kropp, J. R., R. R. Johnson, J. R. Males, and F. N. Owens. 1977. Microbial protein synthesis with low quality roughage rations: isonitrogenous substitution of urea for soybean meal. J. Anim. Sci. 46:837-843. Krysl, L. J., M. E. Branine, A. U. Cheema, M. A. Funk, and M. L. Galyean. 1989. Influence of soybean meal and sorghum grain supplementation on intake, digesta kinetics, ruminal fermentation, site and extent of digestion, and microbial protein synthesis in beef steers grazing blue grama rangeland. J. Anim. Sci. 67:30403051. Lardy, G. P., D. C. Adams, T. J. Klopfenstein, and H. H. Patterson. 2004. Building beef cow nutritional programs with the 1996 NRC beef cattle requirements model. J. Anim. Sci. 82:E83-E92. Lee, G. J., D. W. Hennessy, P. J. Williamson, J. V. Nolan, T. J. Kempton, and R. A. Leng. 1985. Responses to protein meal supplements by lactating beef cattle given a low-quality pasture hay. Aust. J. Agric. Res. 36:729741. Lodman, D. W., M. K. Petersen, C. K. Clark, J. S. Wiley, K. M. Havstad, and M. J. McInerney. 1990. Substitution of DL-methionine for soybean meal as a winter supplement for gestating cows grazing native range. J. Anim. Sci. 68:4361-4375. Loerch, S. C., K. E. McClure, and C. F. Parker. 1985. Effects of number of lambs suckled and supplemental protein source on lactating ewe performance. J. Anim. Sci. 60:6-13. Mackie, R. I. and B. A. White. 1990. Recent advances in rumen microbial ecology and metabolism: Potential impact on nutrient output. J. Dairy Sci. 73:2971-2995. Martin, J. L., K. A. Vonnahme, D. C. Adams, G. P. Lardy, and R. N. Funston. 2007. Effects of dam nutrition on growth and reproductive performance of heifer calves. J. Anim. Sci. 85:841-847. Mathis, C. P., R. C. Cochran, G. L. Stokka, J. S. Heldt, B. C. Woods, and K. C. Olson. 1999. Impacts of increasing amounts of supplemental soybean meal on intake and digestion by beef steers and performance by beef cows consuming low-quality tallgrass-prairie forage. J. Anim. Sci. 77:3156-3162. McCollum, F. T. and M. L. Galyean. 1985. Influence of cottonseed meal supplementation on voluntary intake, rumen fermentation, and rate of passage of prairie hay in beef steers. J. Anim. Sci. 60:570-577. McCollum, F. T., Y. K. Kim, and F. N. Owens. 1987. Influence of supplemental four- and five-carbon volatile fatty acids on forage intake and utilization by beef steers. J. Anim. Sci. 65:1674-1679. Mertens, D. R. and J. R. Loften. 1980. The effect of starch on forage fiber digestion kinetics in vitro. J. Dairy Sci. 63:1437-1446. Mlay, P. S., A. E. Pereka, M. R. Weisbjerg, T. Hvelplund, and J. Madsen. 2003. Digestion and passage kinetics of fibre in mature dairy heifers maintained on poor quality hay as affected by the source and level of nitrogen supplementation. Anim. Feed Sci. Tech. 109:19-33. Momont, P. A., R. J. Pruitt, and P. S. Johnson. 1993. Effect of methionine addition to a urea-grain supplement on intake and digestibility of mature, dormant grasses and performance of cows grazing winter range. J. Anim. Sci. 71:515-521. Ovenell, K. H., K. S. Lusby, G. W. Horn, and R. W. McNew. 1991. Effects of lactational status on forage intake, digestibility, and particulate passage rate of beef cows supplemented with soybean meal, wheat middlings, and corn and soybean meal. J. Anim. Sci. 69:2617-2623. Paterson, J. A., R. L. Belyea, J. P. Bowman, M. S. Kerley, and J. E. Williams. 1994. The impact of forage quality and supplementation regimen on ruminant animal intake and performance. Pages 59-114 In: Forage Quality, Evaluation, and Utilization. G. C. Fahey, Jr., M. Collins, D. R. Mertens, and L. E. Moser (eds) Amer. Soc Agron. Crop Sci. Soc. Amer. Soil Sci. Soc. Amer. Madison, WI. Patterson, H. H., D. C. Adams, T. J. Klopfenstein, R. T. Clark, and B. Teichert. 2003. Supplementation to meet metabolizable protein requirements of primiparous beef heifers: II. Pregnancy and economics. J. Anim. Sci. 81:563–570. 160 Petersen, M. K. 1987. Nitrogen supplementation of grazing livestock. Pages 115-121 In: Proc. Grazing Livest. Nutr. Conf., Univ. of Wyoming. Laramie, WY. Petersen, M. K, D. C. Clanton, and R. Britton. 1985. Influence of protein degradability in range supplements on abomasal nitrogen flow, nitrogen balance and nutrient digestibility. J. Anim. Sci. 60:1324-1329. Reed, J. J., G. P. Lardy, M. L. Bauer, T. C. Gilbery, and J. S. Caton. 2004. Effect of field pea level on intake, digestion, microbial efficiency, ruminal fermentation, and in situ disappearance in beef steers fed forage-based diets. J. Anim. Sci. 82:2185-2192. Reynolds, L. P., P. P. Borowicz, J. S. Caton, K. A. Vonnahme, J. S. Luther, C. J. Hammer, K. R. Maddock Carlin, A. T. Grazul-Bilska, and D. A. Redmer. 2010. Developmental programming: The concept, large animal models, and the key role of utero-placental vascular development. J. Anim. Sci. 88:E61-E72. Rush, I. G. and R. Totusek. 1976. Supplemental value of feed grade biuret and urea-molasses for cows on dry winter grass. J. Anim. Sci. 42:497-505. Schauer, C. S., D. W. Bohnert, D. C. Ganskopp, C. J. Richards, and S. J. Falck. 2005. Influence of protein supplementation frequency on cows consuming lowquality forage: Performance, grazing behavior, and variation in supplement intake. J. Anim. Sci. 83:1715– 1725. Sletmoen-Olson, K. E., J. S. Caton, K. C. Olson, and L. P. Reynolds. 2000. Undegraded intake protein supplementation: I. Effects on forage utilization and performance of periparturient beef cows fed low-quality hay. J. Anim. Sci. 78:449-455. Soder, K. J., V. M. Thomas, R. W. Kott, P. G. Hatfield, and B. Olson. 1995. Influence of energy or protein supplementation during midpregnancy on forage intake of ewes grazing Montana winter range. J. Anim. Sci. 73:2853-2859. Sowell, B. F., J. D. Wallace, E. E. Parker, and G. M. Southward. 1992. Protein supplementation and 48-hour calf removal effects on range cows. J. Range Manage. 45:123-128. Sunvold, G. D., R. C. Cochran, and E. S. Vanzant. 1991. Evaluation of wheat middlings as a supplement for beef cattle consuming dormant bluestem-range forage. J. Anim. Sci. 69:3044-3054. Villalobos, G., D. C. Adams, T. J. Klopfenstein, J. T. Nichols, and J. B. Lamb. 1997a.Grass hay as a supplement for grazing cattle I. Animal performance. J. Range Manage. 50:351-356. Villalobos, G., T. J. Klopfenstein, D. C. Adams, D. Shain, and T. E. Long. 1997b. Grass hay as a supplement for grazing cattle II. Ruminal digesta kinetics. J. Range Manage. 50:357-360. Waterman, R. C., J. E. Sawyer, C. P. Mathis, D. E. Hawkins, G. B. Donart, and M. K. Petersen. 2006. Effects of supplements that contain increasing amounts of metabolizable protein with or without Ca-propionate salt on postpartum interval and nutrient partitioning in young beef cows. J. Anim. Sci. 84:433-446. Wu, G., F. W. Bazer, J. M. Wallace, and T. E. Spencer. 2006. BOARD-INVITED REVIEW: Intrauterine growth retardation: Implications for the animal sciences. J. Anim. Sci. 84:2316-2337. Zhu, X., C. W. Deyoe, K. C. Behnke, and P. A. Seib. 1991. Poured feed blocks using distillery by-products as supplements for ruminants. J. Sci. Food Agric. 54:535. 161 162 1 10.9 DMI, kg 0.30 0.20 Ca, % DM P, % DM 0.29 0.19 Ca, % DM P, % DM 0.21 0.31 903 10.69 9.1 12.6 1.34 59.9 0.21 0.32 848 11.18 9.1 11.4 1.36 60.9 2 8.2 11.5 1.30 58.6 3 0.19 0.29 856 9.92 8.2 12.9 1.25 57.6 0.19 0.30 801 10.38 Table adapted from NRC (1996); MP = metabolizable protein. 1 824 10.10 MP, g/d CP, % DM 7.6 12.2 DMI, kg Milk, kg/day 1.30 58.7 NEm, Mcal/kg TDN, % DM 545-kg mature weight, 9.1 kg peak milk production 769 10.54 MP, g/d CP, % DM 7.6 1.32 NEm, Mcal/kg Milk, kg/day 59.6 TDN, % DM 454-kg mature weight, 9.1 kg peak milk production Item Table 1. Nutrient requirements of mature beef cows1 0.18 0.26 772 9.25 6.6 12.5 1.21 56.2 0.18 0.27 717 9.65 6.5 11.1 1.23 57.0 4 0.17 0.24 688 8.54 4.9 12.0 1.17 54.7 0.17 0.24 633 8.86 4.9 10.7 1.19 55.4 5 0.15 0.22 620 7.92 3.5 11.7 1.12 53.4 0.15 0.22 564 8.17 3.6 10.3 1.14 54.0 6 0.12 0.15 442 5.99 0.0 11.0 0.81 44.9 0.11 0.15 386 5.98 0.0 9.6 0.81 44.9 7 Month since calving 0.12 0.15 454 6.18 0.0 11.0 0.84 45.8 0.11 0.15 397 6.16 0.0 9.5 0.84 45.7 8 0.12 0.15 476 6.50 0.0 10.9 0.90 47.1 0.11 0.15 416 6.47 0.0 9.5 0.88 47.0 9 0.16 0.26 513 7.00 0.0 10.9 0.97 49.3 0.15 0.24 449 6.95 0.0 9.5 0.97 49.1 10 0.16 0.25 574 7.73 0.0 11.0 1.08 52.3 0.15 0.24 502 7.66 0.0 9.5 1.08 52.0 11 0.16 0.25 670 8.78 0.0 11.2 1.21 56.2 0.15 0.24 587 8.67 0.0 9.7 1.19 55.7 12 163 50.2 0.46 7.8 7.16 374 0.22 0.17 50.5 0.46 9.0 7.19 432 0.23 0.18 50.5 0.46 8.8 7.21 422 0.23 0.18 2 50.1 0.46 7.6 7.18 365 0.22 0.17 1 Table adapted from NRC (1996); MP = metabolizable protein. 1 454-kg mature weight TDN, % DM NEm, Mcal/kg DMI, kg CP, % DM MP, g/d Ca, % DM P, % DM 545-kg mature weight TDN, % DM NEm, Mcal/kg DMI, kg CP, % DM MP, g/d Ca, % DM P, % DM Item Table 2. Nutrient requirements of pregnant replacement heifers1 50.7 0.46 9.2 7.18 443 0.22 0.18 50.4 0.46 8.1 7.16 384 0.22 0.17 3 50.9 0.48 9.5 7.22 458 0.22 0.17 50.7 0.46 8.3 7.21 397 0.21 0.17 51.4 0.51 9.8 7.31 476 0.22 0.17 51.3 0.48 8.5 7.32 413 0.21 0.17 Month since conception 4 5 52.3 0.53 10.1 7.52 501 0.21 0.17 52.3 0.53 8.8 7.56 436 0.20 0.16 6 53.8 0.57 10.5 7.89 539 0.31 0.23 54.0 0.57 9.1 7.99 472 0.32 0.23 7 56.2 0.66 10.8 8.53 598 0.31 0.22 56.8 0.66 9.4 8.74 527 0.31 0.23 8 59.9 0.77 11.1 9.62 688 0.30 0.22 61.3 0.81 9.68 10.20 610 0.31 0.22 9 164 10.0 10.6 10.3 0.91 1.38 9.8 1.38 0.33 10.0 0.91 9.3 1.38 9.5 9.5 0.91 0.33 9.0 8.7 1.38 0.33 9.0 0.91 8.2 1.38 8.5 8.4 0.91 0.33 DMI, kg/d 8.0 ADG, kg/d 0.33 70 60 50 70 60 50 70 60 50 70 60 50 70 60 TDN, % DM 50 Table adapted from NRC (1996); MP = metabolizable protein. 1 409 382 354 327 BW, kg 300 1.67 1.34 0.99 1.67 1.34 0.99 1.67 1.34 0.99 1.67 1.34 0.99 1.67 1.34 NEm, Mcal/kg 0.99 1.06 0.77 0.44 1.06 0.77 0.44 1.06 0.77 0.44 1.06 0.77 0.44 1.06 0.77 NEg, Mcal/kg 0.44 Table 3. Nutrient requirements for growing steers and heifers (mature BW = 545 kg)1 10.2 8.4 6.6 10.8 8.8 6.8 11.4 9.2 6.9 12.2 9.7 7.1 13.0 10.2 CP, % DM 7.3 708 594 443 711 589 430 713 584 416 702 570 399 682 550 MP, g/d 379 0.37 0.28 0.19 0.39 0.30 0.20 0.42 0.32 0.20 0.45 0.34 0.21 0.49 0.36 Ca, % DM 0.22 0.19 0.16 0.12 0.20 0.16 0.13 0.21 0.17 0.13 0.23 0.18 0.13 0.24 0.19 P, % DM 0.13 Table 4. Effect of advancing season on nutrient composition of native range in southwestern North Dakota1 Mid-June Late July Early September Early October MidNovember MidDecember NDF, % 59.5 51.0 58.7 59.6 67.9 72.1 ADF, % 35.7 34.8 40.3 38.5 40.9 41.8 CP, % 13.6 14.9 10.2 9.7 6.6 6.2 IVOMD, % 68.1 60.5 55.6 54.3 57.3 53.3 Item 1 Adapted from Johnson et al. (1998); IVOMD = in vitro OM disappearance. 165 166 Primary Primary APR MAY 2 1 1 3 2 3 3 1 1 1 1 # OBS 8.1 ± 0.31 9.1 14.9 15.6 ± 1.73 17.2 ± 3.58 12.4 ± 2.74 17.3 ± 4.58 17.6 29.4 16.6 12.1 CP, % 0.47 ± 0.01 0.48 0.68 0.67 ± 0.07 0.63 ± 0.18 0.67 ± 0.09 0.75 ± 0.13 1.09 0.92 0.91 0.76 NDIN, % 0.19 ± 0.02 0.23 0.14 0.14 ± 0.03 0.08 ± 0.03 0.13 ± 0.04 0.13 ± 0.01 0.17 0.03 0.22 0.12 ADIN, % 1.0 ± 0.06 1.2 1.3 1.8 ± 0.25 2.3 ± 0.78 2.0 ± 0.20 2.3 ± 0.41 3.1 4.2 1.2 1.4 RUP, % 7.0 ± 0.25 7.9 13.6 13.9 ± 1.48 14.8 ± 2.81 10.4 ± 2.56 15.0 ± 4.28 14.5 25.3 15.4 10.7 RDP, % 83.1 ± 0.65 78.6 63.8 70.52 ± 5.01 63.2 ± 11.56 72.0 ± 4.69 68.6 ± 11.99 73.0 49.3 66.8 77.2 NDF, % 55.8 ± 0.98 53.8 44.1 43.7 ± 4.61 44.4 ± 4.29 41.7 ± 4.65 39.2 ± 6.17 42.3 27.3 42.4 50.2 ADF, % 54.2 ± 1.94 47.5 67.7 63.4 ± 4.18 64.4 ± 4.39 66.0 ± 5.28 70.8 ± 5.49 68.3 68.3 60.2 53.4 IVOMD, % Type: Regrowth = growth following July haying; Primary = growth before July haying.; # OBS: number of sampling dates analyzed for a given month, each observation represents 4 to 7 diets collected by esophageal fistulated cows or ruminally cannulated steers; ADIN = acid detergent insoluble nitrogen; RUP = rumen-undegraded protein; RDP = rumen-degraded protein; IVOMD = in vitro organic matter disappearance. 2 Standard deviations listed are for averages of diets collected over 1992 and 1994 within each month, not for laboratory analysis within a particular sample collection. 1 DEC Regrowth Regrowth NOV 2 Regrowth Regrowth Regrowth Primary OCT 2 2 SEPT AUG JUL 2 Primary Primary MAR JUN Regrowth JAN 2 Type Date Table 5. Means and standard deviations of laboratory analysis of meadow diet (OM basis) samples collected in 1992 and 1994 at Gudmundsen Sandhills Laboratory, Whitman, Nebraska1 167 2 2 3 4 3 2 1 2 MAR APR JUN JUL AUG SEPT NOV DEC 6.5 ± 0.60 5.9 7.4 ± 0.34 11.3 ± 2.52 12.3 ± 1.50 13.8 ± 2.53 11.4 ± 1.86 6.0 ± 0.24 6.3 CP, % 0.39 ± 0.08 0.37 0.51 ± 0.06 0.79 ± 0.11 0.90 ± 0.06 0.85 ± 0.15 0.79 ± 0.05 0.48 ± 0.10 0.45 NDIN, % 0.13 ± 0.06 0.27 0.12 ± 0.02 0.16 ± 0.02 0.14 ± 0.01 0.12 ± 0.02 0.11 ± 0.01 0.09 ± 0.05 0.15 ADIN, % 1.2 ± 0.20 0.7 1.1 ± 0.23 1.8 ± 0.43 2.2 ± 0.30 2.5 ± 0.15 1.2 ± 0.11 1.0 ± 0.02 0.8 RUP, % 5.4 ± 0.40 5.2 6.4 ± 0.58 9.5 ± 2.26 10.1 ± 1.31 11.3 ± 2.38 10.2 ± 1.75 5.0 ± 0.27 5.5 RDP, % 86.0 ± 0.96 84.4 79.7 ± 1.30 77.9 ± 4.42 79.8 ± 3.62 72.4 ± 2.70 77.5 ± 5.29 82.5 ± 0.89 83.6 NDF, % 54.5 ± 0.40 56.1 48.8 ± 1.41 46.4 ± 4.28 43.6 ± 4.26 40.6 ± 2.52 43.2 ± 6.17 53.3 ± 0.24 52.5 ADF, % 53.9 ± 5.51 48.3 60.7 ± 1.21 63.7 ± 3.62 67.5 ± 2.4 67.6 ± 2.57 67.6 ± 9.30 54.8 ± 0.69 58.0 IVOMD, % # OBS: number of sampling dates analyzed for a given month, each observation represents 4 to 7 diets collected by esophageal fistulated cows or ruminally cannulated steers; NDIN = neutral detergent insoluble nitrogen; ADIN = acid detergent insoluble nitrogen; RUP = rumen-undegraded protein; RDP = rumendegraded protein; IVOMD = in vitro organic matter disappearance. 2 Standard deviations listed are for averages of diets collected over 1992 and 1994 within each month, not for laboratory analysis within a particular sample collection. 1 1 # OBS JAN Sample date Table 6. Means and standard deviations of laboratory analysis of upland range diet (OM basis) samples collected in 1992 and 1994 at Gudmundsen Sandhills Laboratory, Whitman, Nebraska1,2 STRATEGIC SUPPLEMENTATION TO CORRECT FOR NUTRIENT IMBALANCES G. P. Lardy and R. L. Endecott Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 ECONOMICALLY EFFICIENT SUPPLEMENTAL FEEDING AND THE IMPACT OF NUTRITIONAL DECISIONS ON NET RANCH RETURNS L. Allen Torell1* and Neil R. Rimbey† *Department of Agricultural Economics and Agricultural Business, New Mexico State University, Las Cruces 88005; and †Department of Agricultural Economics and Rural Sociology, University of Idaho, Caldwell Research and Extension Center, Caldwell 83605 ABSTRACT: The difference between biological efficiency and economic efficiency as it relates to supplementation of beef cattle is discussed in this paper. We review well known economic principles and models used to determine the economically optimal supplemental feeding rate for a given cost-price situation. What is known about the economics of feeding supplements for various reasons are explored, including supplemental feeding for improved growth and gain; to maintain and improve reproductive efficiency; to more efficiently harvest rangeland forage and move cattle from sensitive areas by improving livestock distribution; and the importance of supplemental and replacement forages for drought management. The literature review shows that the economics of supplemental feeding is highly variable because beef prices and costs continually change and the output response to supplements is variable and generally poorly quantified. Maintaining high rates of reproductive efficiency is of key economic importance for cow-calf producers and supplementing to maintain the health and condition of the cow herd has significant economic potential, whereas the economics of supplemental feeding to add a few pounds of sale weight is many times not profitable. The economics of supplementation has been shown to be variable and dependent on livestock and rangeland conditions. There is no one prescription or rule that can be followed for making sound economic nutritional decisions. On this basis, we review the basic economic principles and models required to assess economic impacts of supplemental feeding practices and compare the differences between economic and biological efficiency. We review the basic production economic concepts required to assess the economics of supplementation and highlight what is known about the economics of supplemental feeding. Because supplemental feeds have a major role during drought periods we particularly explore the economic impacts of nutritional decisions on ranch returns during dry periods. ECONOMICALLY EFFICIENCT INPUT USE As detailed in any production economic text book (e.g., Workman, 1986), there is a relationship between the amounts of output that can be obtained by different amounts of input use. Numerous shapes of what is called the production function are possible, but the “law of diminishing returns” must eventually hold. By this well known principle, if an input is added while holding other inputs constant, the addition of one more unit of the variable input means output must eventually increase at a decreasing rate and diminishing marginal returns will prevail. For supplemental feeding with a goal of maximizing profit, this means we cannot afford to produce the biggest calf or have the highest possible calf crop. Unless the input is free, the economically optimum level of input use will always be less than that which would give peak biological efficiency, which occurs at the maximum of the production function. The economically efficient or profit maximizing point will occur where the slope of the production function is equal to a linear line that has a slope equal to the ratio of input (supplement) price (Px) divided by output (beef) price (Py). This economically efficient point would occur at X* in Figure 1, Panel A. A different presentation of the economic concepts involved in profit maximization is more common. Recognizing that comparing added value to added cost is the key, Figure 1, panel B shows the standard graphical economic model for optimal input use. It starts with a curve defining the slope of the production function given in Figure 1, what is called marginal physical product (MPP). The MPP curve defines how much more output would be Key words: beef cattle, economic efficiency, feed supplements, rangelands INTRODUCTION Supplementation has been used to improve the growth and rate of gain of stocker cattle and replacement heifers, maintain and improve body condition of cows grazing lowquality forage, improve intake and digestibility of lowquality forage, provide macro- and trace minerals that are deficient in forages, and provide additional nutrients and forage alternatives in periods of drought. Another reason for supplementation that has not been widely explored is the use of supplements to modify grazing distribution patterns on rangelands, and encourage cattle to spend less time in riparian and sensitive areas. Each reason for supplementation can enhance the potential profitability of livestock production. Furthermore, when livestock grazing is tied to sound management that improves rangeland conditions, society may benefit beyond economic returns to livestock producers. 1 Corresponding author: [email protected] 170 replicates. These long-term, multi-rate studies are expensive and rarely conducted over the multiple years and conditions required to estimate the production function that forms the basis of economic evaluations of profit-maximizing input use. Unknown production responses and linkages are the lacking element for evaluating the economics of nutritional decisions. We are left with a partial budget assessment of whether the added economic value of livestock gain and production obtained by feeding at the levels used in selected research trials was enough to justify the added costs. In addition to partial budgeting, in situations where it is established that protein, energy, or some mineral is deficient in the diet, and that a supplement will be fed, calculating the least-cost method of providing for the dietary shortfall may be the best available economic assessment alternative. Evaluating the relative merits of feedstuffs means more than comparing price tags. The $/unit of the limiting nutrient is the important consideration. The price of each available feedstuff potentially used as a supplement is divided by the total pounds of limiting nutrient in a ton of feed, adjusted for moisture content (Torell et al., 1993). Numerous computer programs have been developed to assist in evaluating least cost supplementation and feeding strategies. For example, University of Nevada scientists developed a spreadsheet program to evaluate feeding alfalfa hay to beef cattle (http://www.unce.unr.edu/publications/other/AlfalfaForBee fCows.xls). obtained for each successive increase in input use. Multiplying MPP by Py defines the added economic value, or value of the marginal product (VMP). In a similar way, APP is defined to be the average physical product (APP = Y/X) at each level of X, and when multiplied times Py the average value product (AVP) curve is defined. For a price taking business that can buy as much of the input as desired at the going market rate, marginal factor cost (MFC = Px) is a horizontal line at the specified cost level. Yet, for supplemental feeding decisions, the cost line may not remain fixed at this particular point because wide-spread drought would increase supplement demand, raising input price, and this would increase costs for all producers. The profit maximizing condition is where VMP = MFC (Figure 2, Panel B). The shaded area in the graph defines the profit that would be made, and profit is maximized at X*. At levels of input use below this point additional profit is possible because VMP > MFC. Beyond X* the added cost of the input exceeds the added value. Diminishing marginal returns begins at the top of the VMP curve and this guarantees that a profit maximizing point exists. The part of VMP that lies below the AVP curve defines the producers demand curve for the input. Rearranging the profit-maiximizing condition, Py × MPP = Px to MPP = shows the relationship between Panel A and Panel B. As input prices increase, less input will optimally be used because the slope of the price line increases (Panel A). Or using Panel B, if drought increased supplemental feed costs, the MFC would shift upwards and thus decrease the amount of supplements optimally used. As output price increases, the economic optimum will move closer to peak possible production. A new technology (e.g., a better supplement or forage of improved quality) will potentially shift the production function up and alter the economically efficient level of input use. Value of the marginal product and AVP would shift upwards as well. Economic efficiency is elusive. It varies depending on prices, costs, and production conditions. The simple economic model defining optimal input use has numerous underlying assumptions, and the well known “marginal principles” apply. It is defined for a price-taking business in both the input and output market. A key assumption is that input use during the current year has no impact on production in future years. The shape and location of the production function will vary for different forage and weather conditions, for different animal conditions, and potentially for different breeds and grazing strategies. Defining economically optimal supplementation rates depends on and is limited by estimation of the key biological relationships defined by the production function. One of the biggest limitations in estimating this key relationship is differences in study objectives. The animal scientist applies 2 or 3 alternative supplementation rates and attempts to measure whether there was a significant difference in observed production between treatments. By comparison, estimating the production function requires numerous feeding rates over a wide range, perhaps with less THE ECONOMICS OF SUPPLEMENTAL FEEDING Supplementation for Growth and Gain There are numerous examples of partial budget assessments applied to supplemental feeding practices and these studies generally show a great deal of variability in gain response to supplementation with both energy and protein (Wallace, 1987). Important considerations for predicting performance responses to supplements include forage availability, digestibility, animal growth stage and production requirements, limiting nutrients, and CP content of available forages (DelCurto et al., 2000). Response variability implies a great deal of uncertainty for livestock producers as to whether supplemental feeding for added gain will be a worthwhile economic endeavor. As further noted by DelCurto et al. (2000) and Wallace (1987), alternate-day, once-weekly, and daily feeding of protein supplements have been shown to yield similar production responses, thus labor costs can be reduced by less frequent supplementation. DelCurto et al. (2000) further notes that studies show less variation in supplementation response when greater quantities are fed less frequently, presumably because of less competition for the available supplement. Because calf growth rates interact seasonally with environmental and forage conditions, decisions about optimal sale dates and sale weights become important (Lambert, 1989; Tronstad and Teegerstrom, 2003). VanTassell et al. (1987) found average calf weights in June and at weaning were related to a combination of variables, 171 Using the desirable moderate 5 to 6 BCS score rating as a comparable base, Wikse et al. (1995) found thin cows with BCS of 4 had 74% less economic value when pregnancy rates, weaning weights, and sale price differences were considered. This resulted in a net loss in revenue of $64/cow once reconditioning (increasing body condition) costs were subtracted. Cows sliding to a BCS of 3 had production of only 48% that of BCS 5 and 6 cows, and this meant a net loss of $124/cow. The poor productivity of thin beef cows makes the economics of reconditioning thin cows extremely favorable. Wikse (1995) estimated the benefit to cost ratio for reconditioning thin pregnant cows was over 2:1 ($1 dollar spent reconditioning returns over $2 in economic value). DelCurto et al. (2000) further suggested that because BW change and body condition can be manipulated more easily during pregnancy than after calving, a general strategy for supplemental feeding beef cows is to target a BCS at calving of at least 5 and avoid significant weight loss between calving and the beginning of the breeding season. Starving through a drought is clearly not an economic option given the significant reductions in production from thin cows. The marginal principles of Figure 1 apply. Research data (Richards et al., 1986; Selk et al., 1988; Wikse et al., 1995) indicated the non-quantified production relationship between BCS and calf crop increases very steeply when cows are at a low nutrition level but then flattens near peak levels. A maximum calf crop is not the most economical (Torell et al., 1982), but economic losses from allowing cows to lose substantial weight and body condition are substantial. including calf age, weather conditions, level of supplements fed the cow herd during the winter, crossbreeding and stocking rates. Tronstad and Teegerstrom (2003) found supplemental feeding compensated for the nearly 0 rate of gain for calves reaching 7 to 12.5 months of age. Those authors also found calves supplemented and carried over for May sale were generally the most profitable in central Arizona, where the climate is dependent upon winter rains for cool season grass growth. For the southeast Arizona region, where summer monsoon rains largely determine the production of warm season grasses, lighter nonsupplemented calves also sold in May were most profitable. On the New Mexico State University Corona Range Livestock Research Center (CRLRC) in central New Mexico, where little gain is anticipated over the winter period, sale of weaned calves in the fall was found to be most profitable (Murugan, 2007). Supplementation costs and seasonal sale price differences are obviously an important factor in decisions about supplemental feeding for enhanced gain and calf retention. Some years supplementing and carrying over calves would be the best strategy, whereas in other years, the cost-price situation will not justify the expense. Supplementing the Cow Herd Evaluating the economics of supplementing the cow herd to improve reproductive success means an expansion of the simple economic model detailed earlier. No longer can it be assumed that supplementation decisions during the current period have no impact on production rates in the future. In fact, supplementing the cow herd is done to improve calf crop and production in the future. There are now numerous interactions of economic importance ranging from the body condition of cows before and after calving, cow death rates, conception rates, calving intervals, weaning weights, cull cow weights, and feedlot performance of calves born to supplemented versus nonsupplemented cows. Waldner et al. (2009) evaluated the pattern of losses of cows and replacement heifers in western Canada and concluded that factors related to cow nutrition accounted for 25% of deaths. Stalker et al. (2007) found cow supplementation strategies and weaning date interacted with offspring in the feedlot and impacted feedlot returns well after calves had left the ranch. Including these interactions in an economic model is a challenge that has never been done, but clearly a dynamic model is appropriate. The importance of adequate pre-calving BCS and BW to obtain acceptable pregnancy rates has been well documented (Richards et al., 1986; Selk et al., 1988; Wikse et al., 1995). As noted by Torell and Bruce (1998), a cow that cycles and rebreeds 83 days post-calving will deliver a calf near the same date the following year. Postpartum intervals have been shown to exceed this desired interval when BCS is below the moderate (5 to 6) level, on a 9-point scale. Conception percentage is also reduced from desired rates above 90% to the 70% range for thin cows (Houghton et al., 1990), and thin cows wean lighter calves with reduced survival (Wikse, 1995). Supplementing to Improve Rangeland Conditions Some have argued that government feed programs should be eliminated because feed assistance programs encourage livestock producers to buy supplements and keep livestock during dry periods with stocking levels in excess of sustainable amounts for dry-period forage conditions (Holechek and Hess, 1995). However, potential also exists to improve range conditions using supplements and other livestock management tools designed to improve grazing distribution. Herding, fencing, strategic supplementation, altered seasons of grazing use, development of off-stream water, and early weaning are management options considered by Tanaka et al. (2007) in an economic study about improving grazing distribution. The economic assessment was based on cattle distribution research conducted on the Eastern Oregon Agricultural Research Center Hall Ranch. Based on the grazing trials, projections were made about how weight gain, calf crop, and costs and revenues would change by implementing each alternative management strategy. Resources and costs for a typical 300 cow ranch in northeastern Oregon were used to develop a multi-period linear programming model for the analysis. The production linkages and assumptions about expected production changes from alternative management strategies were the obvious weak link in this economic research, as is usually the case. 172 can support about 683 AUY (16.7 ha/AUY) when forage conditions are near average. In the economic study a conservative stocking rate set at 18.5 ha/AUY (90% of average grazing capacity for the CRLRC) was compared to flexible grazing use where animal numbers were adjusted to match forage conditions. Allowed adjustments to drought included maintaining a flexible yearling enterprise along with a base cow herd, and adjusting annual forage use with yearlings; purchasing replacement forage; and reducing animal numbers. The supplementation practices that Holechek and Hess (1995) criticized, where a protein or energy supplement is fed while animals remain on rangeland and continue to remove forage, was not considered in the economic study. Instead, replacement forage was assumed to be outside leased forage or relatively low-quality hay fed to confined animals. A profit-maximizing, multi-period linear programming (LP) model was used to evaluate how alternative drought management strategies affected net discounted ranch returns computed over a 40-year planning period. Onehundred different beef price scenarios were evaluated in the Monte-Carlo study. Beef prices were different for each year of the analysis (4000 total simulations or iterations with the 40 year planning period), but simulated prices recognized the linkage in price between animal classes and typical price cycles. Annual forage production was considered to be variable in the model based on the finding that herbaceous production on the CRLRC is normally distributed with an average of 656 kg/ha and a standard deviation of 200. “Proper grazing” was defined using guidelines detailed by Bement (1969) and this required that 336 kg/ha of herbaceous plant material remained at the end of the grazing season during a normal or above-average production year. The required residual of herbage was reduced to 224 kg/ha during drought years, which were defined to occur when production was more than 1 standard deviation below the mean (< 451 kg/ha). During dry years, an 11% reduction in sale weights was included and next year’s calf crop was reduced by 5% based on research by Bement (1969), Hart et al., (1988b) and Wikse et al. (1995). As noted by Bement (1969), during periods of drought a livestock producer may decide to leave animals in a pasture even though less than 200 kg/ha of herbaceous production remains. Based on the thin-cow research of Wikse et al. (1995) and others, further reductions in animal performance would be expected as would impacts to future rangeland productivity. The LP model considered the stocking rate restriction to be a maximum amount, penalizing livestock production rates during dry years, but allowing downward herd flexibility if economically optimal. Overgrazing (removing herbaceous production to below 224 kg/ha) was not allowed and animal forage requirements had to be met in all years by either reducing numbers or purchasing alternative feeds. The profit-maximizing model determined the optimal (profit maximizing) choice between selling cows and buying supplemental feed to offset forage shortfalls. It did not determine an optimal level of supplemental feed to use given the complexities and lack of data defining key production relationships. The benefit-to-cost assessment of strategic supplementation assumed that low-moisture blocks would be fed per animal at a rate of 0.32 kg/d for 3 months in the late fall (Oct. through Dec.). The blocks would be placed in previously underutilized areas, and by utilizing these areas, the rangeland grazing period would be extended by a month, reducing hay feeding costs. Grazing costs for the supplement were assumed to increase by $0.23/d or by $6.83/AUM. The extended grazing period would mean 95 tons less hay would be required for the 300 cow ranch. This would mean a $14.59/cow benefit and a benefit to cost ratio of 1.8:1 for the strategic supplementation practice. An important conclusion drawn from the grazing distribution economic assessment was that the most valuable forage is not in June when animals are gaining at peak levels. Instead, the most valuable forage is in the winter or early spring when the forage is dry and low-quality, but few feed alternatives exist. When an expensive purchased feed is the next best alternative, strategies that improve the distribution of animals and more fully utilize less expensive grazing resources can be very profitable. Drought Management Livestock producers have generally endured drought by reducing livestock numbers, leasing forage, temporarily grazing rangelands beyond their capacities, and (or) increasing supplemental feeding. The need for feed supplements and the purchase of alternative forages and hay interacts with the level of stocking; the higher the stocking rate, the higher the likelihood that alternative forages will have to be purchased, or cows sold, when dry conditions occur. Unfortunately, proper grazing practices are not always followed and dry years are when cows become thin with reduced production and lengthened calving intervals. Commonly recommended strategies for dealing with drought and variable forage conditions include maintaining a conservative stocking rate, maintaining grazing flexibility by having yearlings or stocker animals as one of multiple enterprises on the ranch, and leaving a significant amount of herbaceous production at the end of the grazing season (Hart and Carpenter, 2005). In addition to questions about the base level of stocking that should be applied when annual forage conditions are variable, herd mix is another important consideration as different animal classes offer different degrees of flexibility. A common recommendation has been that ranges should be stocked with a basic breeding herd that is not detrimental to the range during drought years, and any excess forage produced in average or above-average years should be utilized by purchased stocker animals. Hart and Carpenter (2005) recommended that in areas prone to drought, breeding herds should constitute no more than 50% to 70% of the total carrying capacity of the ranch during average production years. The rest of the herd should be carry-over yearlings or purchased stocker animals. An economic evaluation of recommended drought management strategies was recently completed using the production rates, costs and characteristics of the CRLRC in central New Mexico (Murugan, 2007). The 11,381-ha ranch 173 or cows were sold (or both) to adjust in these dry years depended on the production alternatives considered to be available, and beef prices. Attempting to feed at least part of the cow herd through a drought was economically optimal in the majority of situations. When only the cowcalf alternative was included, outside forage would be purchased at $50/AUM in 75% of the dry years, and as an average this forage would maintain 225 AUY. Herd reductions would also occur with the average size of the cow herd optimally reduced by 71 AUY. In the 25% of drought years when it was not profitable to replace forage, the cow herd was reduced to an average of 336 AUY, which is 55% of the conservative capacity defined in the analysis. Beef prices were about 3% below average during dry years when herd reductions occurred, but the major consideration was that many of these years were those when drought persisted over several consecutive years. This confirms the obvious. If a prolonged drought occurs, the best strategy is to sell cows. If a flexible yearling enterprise was included, forage leasing was reduced to 22% of the years. The optimal adjustment in this case would be to gradually reduce cow numbers and switch to more flexible yearling production as dry conditions forced cow herd reductions. Yearlings were in the optimal mix, not necessarily because they were more profitable but because in a dynamic situation, they provide a relatively low-cost method to adjust to variable forage conditions. The biggest economic limitation of a conservative stocking strategy is that much forage remains unused during favorable years, and that forage only has economic value to the profit maximizing rancher when it is harvested. As noted by Holechek et al. (2004), implementing a flexible stocking strategy that keeps a relatively constant ratio between forage and livestock is the desire of many rangeland managers, but anticipating and accurately predicting annual forage availability is an impossibility. This limits the practical adoption of flexible stocking strategies. Yet, flexible grazing strategies have significant economic potential to increase ranch profits if accurate projections of forage conditions can be made. Taking advantage of favorable forage production years increased average net annual returns to $69,520 compared to $55,126 with a conservative maximum stocking rate imposed (a 26% increase). Even in the absence of accurate forage and weather forecasts, grazing management can be vastly improved by better incorporating any existing understanding of climate variability into stocking decisions. It is clearly not profitable to starve through a drought (Wikse et al., 1995). Maintaining grazing flexibility and adjusting to forage conditions are the logical economic strategies to follow. It was recognized in the study that few opportunities may exist to purchase outside forage on an as-needed basis during drought, because drought and forage shortages are generally widespread. Two different price situations for leased forage ($20 and $50/AUM) were considered. The lower price reflects a normal full-price forage lease where local forage can be leased with full care provided for displaced animals, or when economically optimal for herd expansion. The $20/AUM price was used during normal or favorable precipitation conditions. The $50/AUM rate reflects the situation where displaced cattle would be shipped a considerable distance or fed relatively lowquality roughage during drought. With variable beef prices and forage conditions, negative net annual returns occurred 22% of the time in the simulation. These poor economic years were largely during short-forage years. When herbaceous production was one standard deviation below the mean (< 451 kg/ha) average estimated net returns (the residual amount left to pay a return to capital investment and to provide a return for management and risk) were -$68,742 compared with +$75,248 during other years. Reduced economic returns carried over for several years with reduced herd sizes and the cost of rebuilding the herd. A key finding of the economic study was the economic importance of following standard drought management recommendations of keeping stocking flexible (Hart and Carpenter, 2005). Herd expansion must largely occur with yearlings to be profitable; increasing and decreasing cow numbers to match forage conditions is too expensive relative to the potential short-term gain, and this is true even when forage conditions are known with certainty. Compared to a cow-calf ranch, adding flexible yearling enterprises increased net ranch returns by 26% with conservative stocking and by up to 66% if an accurate weather forecast made a flexible grazing strategy possible. Without annual variation in forage production, about 84% of available forage would be optimally allocated to cowcalf production. As forage variability increased to levels typically observed on western rangelands, a 50:50 forage allocation between cow-calf and yearling enterprises was found to be optimal. Annual forage variability was an important economic factor. Net annual returns increased from $63,076 to $101,006, a 60% increase, when the economic model was defined without annual forage variability. Leased forage was an economically important strategy for adjusting to drought and for herd expansion during favorable beef price years. With no upper stocking rate restriction imposed, and with both cow-calf and yearling enterprises on the ranch, outside forage was leased in 44% of the years. When a conservative maximum stocking rate set at 90% of capacity was imposed, with cow-calf production only, forage was leased in one-third of the years depending on prices and forage conditions. Of the 4,000 beef price and forage situations considered in the economic analysis, 559 years (14%) were defined to be drought years, based on randomly simulated rainfall and resulting herbaceous production falling 1 standard deviation below the mean level. Whether expensive forage was leased IMPROVING ECONOMIC NUTRITIONAL DECISIONS For the most part, livestock producers must purchase supplemental feeds and make nutritional decisions based on little more than the assurance of the feed salesman that the right economic decision is being made. Beef prices can 174 Hart, R. H., J. W. Waggoner, Jr., T. G. Dunn, C. C. Kaltenbach, and L. D. Adams. 1988b. Optimal stocking rate for cow-calf enterprises on native range and complementary improved pastures. J. Range Manage. 41:435-441. Holechek, J., R. D. Pieper, and C. H. Herbel. 2004. Range Management: Principles and Practices. Prentice Hall.Upper Saddle River, NJ. Holechek, J. L. and K. Hess, Jr. 1995. Government policy influences on rangeland conditions in the united states: A case example. Environ. Monit. Assess. 37:179-187. Houghton, P. L., R. P. Lemenager, L. A. Horstman, K. S. Hendrix, and G. E. Moss. 1990. Effects of body composition, pre- and postpartum energy level and early weaning on reproductive performance of beef cows and preweaning calf gain. J. Anim. Sci. 68:1438-1446. Lambert, D. K. 1989. Calf retention and production decisions over time. West. J. Agric. Econ. 14:9-19. Manley, W. A., R. H. Hart, M. J. Samuel, M. A. Smith, J. W. Waggoner, Jr., and J. T. Manley. 1997. Vegetation, cattle, and economic responses to grazing strategies and pressures. J. Range Manage. 50:638-646. Murugan, S. 2007. Profit maximizing livestock production and marketing strategies to manage climate variability. Thesis. New Mexico State Univ. Las Cruces, NM. Richards, M. W., J. C. Spitzer, and M. B. Warner. 1986. Effect of varying levels of postpartum nutrition and body condition at calving on subsequent reproductive performance in beef cattle. J. Anim. Sci. 62:300-306. Selk, G. E., R. P. Wettemann, K. S. Lusby, J. W. Oltjen, S. L. Mobley, R. J. Rasby, and J. C. Garmendia. 1988. Relationships among weight change, body condition and reproductive performance of range beef cows. J. Anim. Sci. 66:3153-3159. Stalker, L. A., L. A. Ciminski, D. C. Adams, T. J. Klopfenstein, and R. T. Clark. 2007. Effects of weaning date and prepartum protein supplementation on cow performance and calf growth. Rangeland Ecol. Manage. 60:578-587. Tanaka, J. A., N. R. Rimbey, L. A. Torell, D. T. Taylor, D. Bailey, T. DelCurto, K. Walburger, and B. Welling. 2007. Grazing distribution: The quest for the silver bullet. Rangelands 29:38-46. Torell, L. A., C. F. Speth, and C. T. K. Ching. 1982. Effect of calf crop on net income of a nevada range cattle operation. J. Range Manage. 35:519-521. Torell, R., J. Balliette, and L. J. Krysl. 1993. Pricing protein and energy supplements. Beef Cattle Handbook BCH-5451. Available at: http://catalog2.nmsu.edu:2048/login?url=http://sea rch.ebscohost.com/login.aspx?direct=true&db=agr &AN=IND20448659&site=ehost-live&scope=site. Torell, R. and B. Bruce. 1998. Backing up calving dates (reducing postpartum interval of beef cows). Univ. of Nevada - Reno Fact Sheet 98-19. generally be reasonably estimated. The cost of feedstuffs is known when purchased. Thus, economic variables are not the key factors limiting the economic decision. Instead, it is unknown response and production relationships (Figure 1) that primarily limits sound nutritional decisions. Long-term studies that evaluate animal responses to grazing systems, stocking rates, and alternative levels of supplementation, like those conducted by Hart et al. (1988a) and Manley et al. (1997), are now rarely funded. The emphasis of agricultural research and extension programs has changed greatly as we have evolved to a more urban and environmentally aware population. Colleges of Agriculture and traditional range and animal science departments have added the terms “environmental”, “ecology”, and “conservation” to their names. Funding priorities have changed and the new priorities are not the long term grazing studies needed to quantify key production relationships required for economic evaluation of nutrition-based grazing alternatives. Within the current research environment, the key to improving the ability of grazing and feeding trial research to contribute to improved economic decisions is to remember that more and varied rates of study treatments are important. Furthermore, when the top input rates of the study begin to cause output to increase at a decreasing rate, it does not mean that the maximum relevant rate has been reached. Instead, it means the beginning of the economic relevant production stage has been reached. Measurement over a number of years with varied conditions is important. Partial budgeting of a limited number of feeding rates will likely remain the key analysis tool given research budgets and priorities. Several key points about the economics of supplemental feeding practices are clear from the view of a couple of economists. First, the marginal principles very much apply to nutritional decisions. Just because some new feed supplement or additive increases weaning weights or calf crop does not mean it is economically justified. One must consider the added cost and added value of adopting the practice. Further, there is an economically optimal rate of the production input to use and it will be less than that which would yield the biggest and most output. LITERATURE CITED Bement, R. E. 1969. A stocking-rate guide for beef production on blue-grama range. J. Range Manage. 22:83-86. DelCurto, T., B. W. Hess, J. E. Huston, and K. C. Olson. 2000. Optimum supplementation strategies for beef cattle consuming low-quality roughages in the western united states. J. Anim. Sci. 77:1-16. Hart, C. R. and B. B. Carpenter. 2005. Stocking rate and grazing management. Texas Agric. Ext. Serv. E64. Hart, R. H., M. J. Samuel, P. S. Test, and M. A. Smith. 1988a. Cattle, vegetation, and economic responses to grazing systems and grazing pressure. J. Range Manage. 41:282-286. 175 Tronstad, R. and T. Teegerstrom. 2003. Economics of sale weight, herd size, supplementation, and seasonal factors. J. Range Manage. 56:425-431. VanTassell, L. W., R. K. Heitschmidt, and J. R. Conner. 1987. Modeling variation in range calf growth under conditions of environmental uncertainty. J. Range Manage. 40:310-314. Waldner, C. L., E. G. Clark, L. Rosengren, and R. I. Kennedy. 2009. A field study of culling and mortality in beef cows from western canada. Can. Vet. J. 50:491-499. Wallace, J. D. 1987. Supplemental feeding options to improve livestock efficiency on rangelands. Pages 92-100 In: Achieving Efficient Use of Rangeland Resources. R. S. White and R. E. Short (ed) The Fort Keogh Res. Symp. Sept., 1987. Miles City, MT. Montana State Univ. Agric. Exp. Sta. Bozeman, MT. Wikse, S. E. 1995. Use of performance ratios to calculate the economic impact of thin cows in beef cattle herds. N. Amer. Vet. Conf. Proc. 9:507-508. Wikse, S. E., D. B. Herd, R. W. Field, P. S. Holland, J. M. McGrann, J. A. Thompson, C. White, and R. Angerstein. 1995. Use of performance ratios to calculate the economic impact of thin cows in a beef cattle herd. J. Amer. Vet. Med. Assoc. 207:1292-1297. Workman, J. P. 1986. Range Economics. McMillan Publishing Co. New York, NY. 176 Panel A Panel B Added Value and Cost ($) 60 VMP 50 40 AVP 30 Profit MFC 20 10 0 0.000 5.000 10.000 15.000 20.000X* 25.000 30.000 Level of Input use (X) Figure 1. Graphical presentation of the traditional economic model of optimal input use. 177 ECONOMICALLY EFFICIENT SUPPLEMENTAL FEEDING AND THE IMPACT OF NUTRITIONAL DECISIONS ON NET RANCH RETURNS L. Allen Torell and Neil R. Rimbey Notes Proceedings, Grazing Livestock Nutrition Conference July 9 and 10, 2010 CHALLENGES TO PREDICTING PRODUCTIVITY OF GRAZING RUMINANTS: WHERE TO NOW?1 M. K. Petersen2ŧ, J. T. Mulliniks*, A. J. Roberts ŧ, and R. C. Watermanŧ ŧ USDA - ARS, Fort Keogh Livestock and Range Research Laboratory, Miles City, MT 59301; Department of Animal Science and Range Sciences, New Mexico State University, Las Cruces 88003 * ABSTRACT: The Fourth Grazing Livestock Nutrition Conference was convened at Estes Park July 9 and 10. There were over 28 poster presentations and 12 conference papers presented. The papers were organized in 6 topical sessions ranging from microbiology to supplementation. The first session covered the potential for rumen microbiological studies in grazing animals. Then the keynote address focused on methods to assess intake of grazed vegetation by the use of alkane as an internal and external marker. Followed was a discussion of the methodology and types of analysis for assessment of grazing animal behavior, which is considered a key component of livestock distribution. There was a very informative session covering the metabolic and practical consequences of weight loss, undernutrition during pregnancy, and nutrient excess associated with high-quality forage diets. Integration of the previous session into discussion of prediction models and their value to predict animal productivity in Australia and US followed. The conference finished with a summary of the state of the art in regards to supplementation strategies and the implications of nutritional decisions on net returns. The conference provided current information and inspired a number of new research ideas by the participants. advances to keep pace with the need for new information, the authors suggested the importance of grazing livestock nutrition is greater than before and requires more emphasis. To keep the rate of advances there needs to be a rebirth of research endeavors and conservative funding of the past is no longer appropriate. In another context, the authors effectively pointed out an ongoing dilemma in accurately assessing intake and described successes in both research and practical arenas in the elucidation of supplementation effectiveness and management. Records of these advances can be found in conference proceedings of the first, second, and third meetings. Authors Tim DelCurto and Ken Olson linked research success to production and economic application as the ultimate evaluation of research outcomes. They pointed out the contradiction with Journals preferring mechanistic results when the setting for grazing livestock nutrition research is the field environment. Their statement implies that a number of important applied studies have not been published. RUMEN MICROBIAL ECOLOGY AS INFLUENCED BY THE NUTRITIONAL ENVIRONMENT OF GRAZING RUMINANTS Presentations on methods and procedures to manipulate microbial populations in grazing animals organized by Bryan White (Brulc et al., 2010) and Shanna Ivey (LodgeIvey, 2010), illustrated prospective impacts this area of research may have on grazing livestock nutrition. Riddled with a number of challenges, this area is largely unexplored. One of the major success stories highlighted was introduction of bacteria into Australian goats that are capable of degrading 3-190 hydroxy-4 (1H)-pyridone (DHP). The Australian animals were then able to use Leucaena leucocephala, which contained the toxicant mimosine that degrades to a toxic goitrogen. Other promising developments were presented and evaluated. The future is wide open for researchers in the study of rumen microbiology in free ranging ruminants since the field is young. For example, a basic evaluation of the microflora nutrient needs in relation to nutrients consumed by the host is not well defined for native vegetation diets. Knowledge of microflora nutrient limitations under grazing conditions will potentiate the ability to augment nutrient supply. INTRODUCTION This ambitious conference, arranged jointly by Western Multistate Research Project W-1012, the University of Wyoming, and Colorado State University brought together a number of experts in grazing livestock nutrition. The conference was comprised of presentations from domestic and internationally renowned researchers. It began with an excellent overview of Issues in Grazing Livestock Nutrition, which framed the conference agenda. The authors (DelCurto and Olson, 2010) first described hallmarks in progress attained by past and current researchers. For future 1 USDA, Agricultural Research Service, Northern Plains Area, is an equal opportunity/affirmative action employer, and all agency services are available without discrimination. Mention of any trade name or proprietary product does not constitute a guarantee or warranty of the product by USDA, Montana Agric. Exp. Stn., or the authors and does not imply its approval to the exclusion of other products that also may be suitable. 2 Corresponding author: [email protected] 180 interferes with cross-location comparisons, outlook in interpretation, and publication of results. A sidebar comment by Anderson (2010) suggested the most effective method of grazing management to best utilize high-quality vegetation was with a herder. The application of technology (e.g., a virtual fence) to move (herd) stock to areas desired without the expense for labor should be commercially available soon. Both Anderson (2010) and DelCurto and Olson (2010) pointed out the importance of understanding free-ranging animal behavior and the deficit in this knowledge is slowing advancement of understanding animal distribution and grazing animal nutrition. Anderson (2010) also shared postulates that can be used to structure and stratify behavioral data for interpretation and analysis. He suggested that single points of observations (i.e., 1 animal) could be misleading at best, because it ignores the gregarious nature of domestic livestock and their interdependence. This suggestion was driven home by encouraging experiments with (at a minimum) typical naturally sized social groups. Another challenge across locations is to integrate dissimilarities among instruments, methodologies, and interpretations. Mechanisms guiding animal distribution (imbedded within nutrient location richness) are the least understood characteristic of grazing animal nutrition that may be accounting for a large portion of nutrient consumption. He continued by suggesting foraging decisions are the balance between the motivation to eat and satiation (which do not occur in isolation). We have more to discover before we can successfully integrate animal landscape use as part of a prediction tool for nutrient capture by grazing animals. PERSPECTIVES ON INTAKE IN GRAZING STUDIES AND IN PRACTICE The next section of the conference focused on assessment of the use of landscapes by the grazing ruminant. The keynote talk was presented by Hugh Dove. Dr. Dove addressed the research application of alkane assisted simultaneous digestibility and fecal output measurements (adoption and evaluation of the alkane technique is an objective of W1012). First, Dove (2010) described the historical slow progress in perfecting methodology for measurement of grazed vegetation consumption. He went on to explain the thought process behind the alkane technology and the possibility that if the assumptions were true then this technology would be a breakthrough for our discipline. This methodology has been slow in adoption by researchers in the western US. With the loss of the slow release bolus, intake assessment is less convenient. This is reflected by fewer intake measurements in recent literature, making those reports less complete and resulting in an unfortunate gap in knowledge. The promising aspects according to Dove (2010), suggest that the technique does not require quantitative evaluation of alkanes in grazing animal diets as long as the ratio of plant (diet) derived alkane in the feces and that of a dosed similar chain length alkane can be determined. Support for this important assumption of the technique is dependent on nearly equal total tract digestibilities of 2 similar length alkanes. Therefore, the fecal ratio should be equal to the feed ratio. This assumption is a central tenet to move towards implementation of the method, especially for workers conducting livestock intake studies on extensive native rangelands. Another benefit of this dual marker system is that only 1 laboratory analysis needs to be conducted not 2 separate analyses for each component of the duo-markers. It is clear that the challenge of the alkane technique is assessment of its capability and robustness in specific experimental settings. Dove (2010) pointed out a number of appealing attributes, such as: the technique adjusts to each animals digestive character; the external marker is required for determination of both digestibility and fecal output; only 1 laboratory analysis is conducted; measurement can be conducted with animal fed supplements; relative concentrations are important (not quantitative evaluation); and accuracy is less susceptible to variation caused by changes in physiological state. If researchers achieve accurate results by applying this technique, then we will have overcome a major obstacle to predicting productivity of grazing ruminants. The second paper in this session discussed measurement peripheral to intake assessments such as grazing behaviors. Dean Anderson explained how new geospatial technologies pioneered first by the military and adapted by wildlife researchers are being used to understand landscape use by domestic ruminants. The data recording potential of instruments can collect a datum point every second, which leads to precise differentiation of location, rate of movement, and potential identification of specific behaviors. Not having a universal model to analyze and interpret all location data is perhaps realistic but problematic. However, the lack of continuity probably IMPACT OF NUTRIENT BALANCE ON BIOLOGICAL EFFICIENCY AND LIFETIME PRODUCTIVITY OF GRAZING LIVESTOCK Adaptive metabolic changes occurring in beef cattle when the nutritional or external environment becomes extreme was the focus of one of the afternoon sessions. The first example began with the paper by Richard Waterman and Ron Butler, who contrasted metabolic dysfunction of the high producing dairy cow with the postpartum range cow. These scenarios are similar metabolically but they have different causations. Dairy cows are prolific milk producer fed a very high-quality diet whereas range beef cows produce much less milk and consume low-quality diets. The net effect of each circumstance is that nutrient intake is insufficient to meet nutrient demands for milk production regardless of cow type, which results in BW loss. The dairy literature possesses a deep body of work revealing the mechanisms at the root of metabolic dysfunction. Few range cow studies have the stepwise methodology characteristic of the efforts in dairy cow research. As an area of investigation, the phenomenon of weight loss and its efficiency lacks definition although it is a predictable event with known negative consequences. Waterman and Butler (2010) made the case that an inadequate supply of glucose maybe one of the causes of metabolic dysfunction. Speculation on the cause of a low supply of glucose included low propionate derived from ruminal fermentation and an increased demand for glucose 181 environments, they make the point that excess nutrients are also a stressor. A key point was that a changing nutrient supply during the peri-conceptual period can have profound impacts on reproduction. Implication of nutritional status in embryo survival is a key factor influencing efficiency of assisted reproductive management. Evaluation of nutritional status of an individual is a broad term encompassing the sum of the current trend in nutrient intake and (or) use as well as the product of historically accumulated impacts occurring from the embryonic stage to a current point in time. The authors lead the reader to make a critical observation, that the nutritional history of females used in experimental evaluations has been neglected as a source of variability that seems to have overwhelming implications on cow responses within a study or when comparing across studies from different laboratories. This conclusion is derived from experimental results showing gene expression differences driving a number of functions, such as adipocyte growth and differentiation, lipogenesis, angiogenesis, and anti-angiogenic operations. They suggest that experimental data show low nutrition during the perinatal period will increase offspring mortality. Additionally, a lower plane of nutrition during lactation may reduce absolute BW of the neonate and throughout the growing period because calves experiencing lower planes of nutrition during the first 6 months of life are less apt to compensate later in life. Overall, the consequences of undernutrition as described are variable depending on the severity, timing, and the pre and post nutrient intake in relation to an undernutrition event. In some scenarios, the reader could construct a case where undernutrition may improve efficiency later or severe undernutrition has profound negative implications. As a consideration, we are challenged to predict short- and long-term influences of undernutrition as impacted by the neonate’s birth weight. This relationship gives an opportunity to exert a little management since there is a degree of control over birth weight via sire selection. Some of the effects described by the authors may be ameliorated by intentional selection of lower birth weight sires to reduce physiological state induced stress. Then the question becomes, Will the impacts of metabolic adaption be significantly modulated by a 5, 10, or 15% decrease in birth weight? It is probable that supply of specific metabolic intermediates is playing a primary role in pregnancy and nutritional status interactions affecting the trends toward weight loss, maintenance, or gain. Lastly, the lifetime consequences of maternal nutritional status on progeny as a neonate and as an adult may have strong influences on traits of economic importance. For the final presentation in the afternoon session pertaining to nutrient balance on biological efficiency, Monty Kerley (Kerley, 2010) asked the question “Do ruminants experience imbalances when they consume high quality forages? His talk started with a reflection on the finding that there is considerable variation in 2 associated measures of animal responses to their diet (i.e., intake and gain). He pointed out that cattle with the same rate of gain can have a 40% difference in the quantity of feed consumed. It was proposed that intake differences relate to divergence in metabolic efficiency among individuals and to support milk production resulting in supply exhaustion. This scenario is associated with body tissue mobilization and weight loss. Mobilization of adipose for utilization in oxidative metabolism and protein tissues degraded to AA with a significant portion of carbon skeleton diversion to gluconeogenesis, while other AA are utilized in oxidative metabolism. Description of metabolic consequences of weight loss, involvement of 6 organs in this process, metabolic foot prints highlighting metabolic adjustments, and general interpretations were presented. A description of changes (or signals) and analytical measurements to verify the severity of nutrient deficiency were suggested. The case was made that knowledge of weight loss and practices to ameliorate weight loss are crude at best. Methods to increase glucose supply in range supplements and application of a hand-held, chute-side ketone meter for rapid assessment of glucose status were discussed. Adaptive changes by a cow to weight loss may alter (hopefully improve) her ability to cope with the external and nutritional environment. These metabolic modifications have an impact on energy metabolism and expenditure and most likely carcass composition and nutrient priority utilization. Proportional fate of absorbed AA for protein tissue anabolism, gluconeogenesis, oxidative metabolism, fetal demand, or milk production is poorly understood during weight loss. The pathway a fraction of metabolic protein follows while cows undergo metabolic adjustments most likely has a profound influence on efficiency and response to dietary nutrients. Alteration of this phenomenon probably is modified by nutritional history, such as: In the recent past, was a cow fat or thin? Is she increasing or decreasing in condition? Has she been grazing range her entire life or was she developed and raised in a feedlot? Alternatively, was the cow developed in a region with 50% greater or lesser precipitation than where she is today? The paucity of knowledge on the impact of these factors interferes with our ability to develop the most efficient nutritional management schemes. Presentation and discussion of the importance of nutrient balance and nutrient sufficiency in the pregnant ewe and cow by Joel Caton, who with Bret Hess, examined implications of dam nutritional status on her progeny’s lifelong vigor. Emphasis was placed on progeny growth, development, and performance outcomes as modified by timing of nutritional perturbations and potential long-term production effects. Discussion started with known conditions that are outcomes of alternative metabolic patterns (i.e., the thrifty phenotype and predictive adaptive responses) that promote immediate well being. The consequences of these metabolic patterns may include insulin resistance, altered shape of growth curve, and body composition alterations at maturity. From their ideas, it is conceivable to propose that slight to moderate maternal nutrient restriction occurs in a high proportion of western range cows. As suggested by Waterman and Bulter (2010), and reiterated by Caton and Hess (2010), there is evidence to support the theory that maternal nutrient supply, physiological demand, and environmental conditions (energy demand) lead to stress responses modulated or amplified by nutrient restriction. Although the major portion of discussion concerned inadequate nutritional 182 the case was made that more efficient animals will be more widespread in the future as a result of preferences for bulls with known lower feed intake will be selected. This practice will most likely affect caloric requirements and simultaneously impact the proportional need and demand for other nutrients. In case of grazing high-quality forages, post-ruminal supply of AA (microbial and feed origin) may not be sufficient to support energy-allowable gain creating a disparity between energy and protein. In fact, he reported that arginine, histidine, leucine, lysine, methionine, and threonine are limiting, with the most limiting being methionine in high-quality forage diets. Therefore, he focused on potential imbalance of absorbable AA to available energy on cow and weaned calf performance consuming a forage diet. He discussed the results of cow selection based on relative feed intake and identified cows that maintained the similar body condition and weight while consuming significantly different amounts of feed. He noted that their daughters followed the same trends with lesser magnitudes in intake differences. Because the AA composition of microbial protein is nearly constant, it is easy to predict sufficiency of protein composition in comparison to available energy. From those predictions, it was determined that a grazed forage intake with sufficient calories to support 0.9 kg/d would be limited by methionine supply equivalent to 0.3 kg/d. He went on to contrast how supplementation with rumen-undegraded protein (RUP) to young lactating cows reduced milk production and improved BW gain while supplementation with additional energy increased milk production and increased BW loss. He concluded that this tool could be effect for managing the challenges young cows face in nutrient allocation between growth and milk production. Conclusion of the discussion occurred with results of studies demonstrating a partial response to the above-described response is partially due to the supply of methionine. Kerley (2010) suggested that 7 g of methionine per day has potential to improve heifer growth and development. The take home message of the presentation was voluntary intake of beef cattle in the future will most likely be less while maintaining at least the same productivity (i.e., high efficiency). Those foreshadowed changes are in proportions and absolute nutrient density of diets for the future. It is very possible that the scenario described above has already occurred in isolated cases where a strict selection criterion has been followed in the development of a cowherd in various environments. The utilization of RUP to partition nutrient use in young cows is a powerful tool to reduce nutritional stress in young cows. In practice, intake responses have not been intentionally identified but efficiency has been indirectly selected for where conservative feeding practices have been exercised with strict reproductive standards imposed on the heifers and cows. In these environments, only cows that are the most efficient will be productive in a nutrient conservative environment. APPLICATION OF REQUIREMENT SYSTEMS FOR GRAZING LIVESTOCK: ASSESSING NRC AND CSIRO SYSTEMS FOR GRAZING LIVESTOCK The second day started with a talk on assessing nutrient requirement schemes for grazing animals. Stuart McLennan and colleagues (Dove et al., 2010) showed the strengths and limitations of how the NRC and CSIRO models can be used to understand the role grazed vegetation plays in animal growth and maintenance. At the outset, those authors demonstrate the difficulty we have is characterizing nutrient intake at a variety of levels and referred back to the presentation addressing alkanes for intake determination. Making the point that no matter how good a prediction models may be, it will not be effective if the predicted nutrient intake is inappropriate. Therefore, a perspective proposed by the authors examined the use of models to predict animal performance or predict forage adequacy. With that in mind, they then make the generalization that current accuracy of published energy requirements seems sufficient for application to nutritional management of grazing livestock (in temperate environments). A weakness of systems is the present inability to account for N recycling, especially in dormant forage settings. They also pointed out that protein demand sometimes interacts with issues, such as parasites, but they felt these relations are not so great that they render the application of the models ineffectual under field conditions. The authors pointed out different approaches accounting for the dynamics of maintenance requirements by the CSIRO and Cornell systems. The Australian system accounts for maintenance energy modifications with changes in intake, whereas the Cornell model adjusts with metabolic size. Scaling of maintenance requirements within the CSIRO system occurs in relation to a standard reference weight. An example of a gap in the requirements includes the utilization of body tissue (weight loss) during pregnancy that is masked by increases in hydration. Finally, the discussion leads to the influences of management goals in how we use prediction models. Key points included available biomass (kg/ha), perspective of productivity per animal in comparison to productivity per ha, acceptability of nutrient imbalances, or reduced nutrient intake and subsequent suboptimal performance. SUPPLEMENTATION STRATEGIES TO ACHIEVE BIOLOGICAL AND ECONOMIC EFFICIENCY The second morning session focused on review of supplementation as a nutritional management tool. The authors, Greg Lardy and Rachel Endecott (Lardy and Endecott, 2010) suggested that the recent rise in feedstuff prices has created a heightened awareness of supplementation and feeding in general as a major cost of grazing animal production. They illustrated the dynamic nature of the supplementation decision process by integrating changing forage quality and shifting nutrient requirements of the female while considering the needs of both the ruminal microflora and the host animal. This complexity is compounded by interactions of physical and chemical form nutrients are supplied in a supplement. They 183 also presented circumstances where supplement used to improve grazing distribution had value due to the improvement in the duration of grazing in a pasture. Lastly, the authors discussed the value of flexibility for dealing with changing conditions using a base cowherd augmented by disposable yearling or stocker cattle. They summarized by encouraging researchers to conduct multiyear experiments in varied conditions and to use partial budgeting as a method to evaluate. pointed out that form of carbohydrate, source of N and potential impact of AA composition, and microbial nutrient and (or) co-factors needs can be important considerations affecting efficiency. Animal responses in general are highly predictable but in some specific situations predictability is much lower. The largest source of variability influencing responses is associated with the inconsistency of the year (precipitation, snow accumulation, duration of high and low temperatures, forage production, etc.). Responses due to nutritional management are specifically unpredictable each year. Lardy and Endecott (2010) suggested milk production would increase with an increased supply of metabolizable protein (MP). They pointed out that supplementation to satisfy MP needs will tend to increase pregnancy rate. The apparent contradiction in the effectiveness of infrequent protein supplementation was noted as was delivery cost saving associated with that management scheme. Another role protein provision can play is to reduce the negative associative effect on digestibility when non-structural carbohydrates are fed and other interactions occurring with carbohydrate sources. The authors finished their discussion on nutrients/feedstuffs with a summary of reviews of literature on research with addition of fat to supplements. They suggested that fat is not generally an effective feed source to increase dietary energy intake. However in some cases, fat may have beneficial effects on reproduction in heifers or cows when fed before calving. They closed their presentation by suggesting that more thorough and better characterization of seasonal changes in degradability of forage CP, as well as forage intake is needed for successful and cost effective supplementation programs. The final talk of the meeting by Torell and Rimbey (2010) proposed integration of the previous topics into an economic analysis. With this purpose, those authors set out to clarify the economic goals of domestic animal grazing and nutrition. For the range beef cow, maintaining high rates of reproductive efficiency is a key economic driver. Supplementation to augment economic success can occur by improving health or condition of brood cows, but to add a few kg to sale weight via supplementation is not likely to give a feasible return. There are an infinite number of outcomes when evaluating the amount of output in comparison to different amounts of input. This idea was framed by the comment that added value (due to supplementation) should be compared to added cost. The predictability of the net effect is confounded by the elusive nature of economic efficiency, which varies by price and production conditions. Publications of experiments to evaluate manipulation of cow metabolic efficiency driven by inputs using an economic outcome as the response criteria are rare. Few long-term, multi-rate studies and multiyear studies are conducted. However, multiple year studies are required to estimate the production function that forms the basis of economic evaluation. There is a need for long-term studies to evaluate practices that can suggest alternatives for future production practices. This concept is illustrated by the depth in the body of work used to assign economic values to cows with various BCS. At a BCS of 4, the author’s estimated 74% less economic value compared to a cow with a BCS of 5. In this context, 1 dollar spent in reconditioning returns 2 dollars in economic value. They EMERGING CONCEPTS IN GRAZING LIVESTOCK NUTRITION – WHERE TO NOW? The various grazing environments described in the presentations ranged geographically from Missouri to North Dakota to New Mexico to Oregon to Australia. The common thread among the studies described in these papers is a grazing ruminant harvesting plant materials, but the environments are different. Despite obvious variability in locations where grazing cattle are managed, the paper by Dove et al. (2010) comparing prediction models was quite promising. In some conditions, the models predict animal responses well. In other presentations, we were made aware of situations that can lead to temporary or long-term changes in animals’ responses to environment and nutrition; these situations may not fit the models as well. Although research has been conducted under this assumption, expecting linear response in a physiological system reflects lack of knowledge that responses to supplemental nutrients are linear. We know that responses to limiting nutrients can be curvilinear, as described by the economic diminishing returns curve discussed by Torell and Rimbey (2010). In fact, the classical methodology for determination of nutritional requirements uses a dose response evaluation. When the response ceases with incremental increases in dose then that point is determined to be the requirement. At the Corona Range and Livestock Research Center in New Mexico, cows have been developed and maintained over the past 10+ years using a conservative nutritional management scheme. This site would be considered a piñon-juniper rangeland with blue grama as the predominate grass. Normal yearly precipitation is 330 mm. Replacement yearling heifers achieve 50 to 55% of their mature weight by breeding. Over the last 8 years, pregnancy rate in heifers has averaged 80% in a 45-d or less breeding season (Table 1). With this type of heifer development where 20% of heifers are culled for reproductive failure, it is assumed that the fittest heifers become pregnant (Hawkins et al., 2000). Subsequent pregnancy rates for these females has averaged 92% as 2year-old cows after their first calf with a 60-d breeding season, 90% as 3-year-old cows, and 95% as 5-year-old and older cows. In most years, all cows receive less than 100 kg of a 36% CP supplement per year with the remainder of their consumed nutrients supplied by grazed native range vegetation (Mulliniks et al., 2009). The above described management scheme inspired us to begin investigating methods to achieve optimal biological and economic efficiency. We set out with a series of studies to inspect the lower amounts of protein supplement needed 184 to elicit highly efficient optimal responses. To assess efficiency, we compared our animal responses to the predicted gain from the NRC Beef Cattle Requirements. The first scenario to be presented will be cows 5 years old and older, grazing dormant range in the winter (grazable forage was not limiting; Table 2). The cows were bred to begin calving the third week in February. Their initial BCS in October at weaning were 5.0 + 0.1. Over the ensuing 4year period, duration of supplementation was varied depending upon conditions of the year. Duration of supplementation ranged from 25 to 90 d (average 61 d). Cows received 1 of 3 supplement treatments: 3.2 kg per week of a 36% CP cottonseed meal-based supplement fed in equal amounts on Monday, Wednesday, and Friday (CSM), 1.6 kg per week of a 65% CP (or higher) animal protein-based supplement (either blood and feather meal or fish meal) that was self-fed with a mineral mix (SMP), or a negative control where cows received the CSM at 0.3 kg per week as prescribed at the discretion of the manager depending upon weather conditions (NC). The 3 supplement regimes supplied 72, 49 or 45% of the CP requirement (for CSM, SMP, and NC, respectively). Actual vs. predicted weight change for the 61-d period was -0.2 vs. -54; 1.8 vs. -62.8; and -12.6 vs. -64.7 kg/d for CSM, SMP, and NC, respectively (Sawyer et al., 2010). There are 2 thoughts suggested from these cow responses. The efficiency at which CP was utilized varied greatly between the CSM and the SMP supplement. The response to supplement is confounded, however, by the delivery method since CSM was hand-fed 3 days per week and the SMP supplement was self-fed and always available. The 2 supplements varied in ruminal degradability. The second observation is that the actual and predicted responses were not in good agreement. The explanation for this lack of agreement could be due to the inaccuracy of the values we used in the model or our cows were different than the cows whose data were used to develop the model. We propose that our cows are highly adapted to their environment and are therefore more efficient (Heilbronn et al., 2006). The first indication that these cows have the potential to be more efficient is related to their amount of activity. Cibils (2009, personal communication) described distance traveled by cows will be in excess of 16 km/d even when water is within 0.5 km. When reviewing literature investigating the responses to exertion, a few key responses stand out. Heart rate declines, oxygen consumption declines, cellular mitochondria increases, and insulin sensitive increases, all which allow for greater energetic efficiency in spite of increased expenditure. A second possible explanation is a potential reduction in organ mass compared with traditionally managed animals. Moderate energy restriction has been shown to decrease organ mass and maintenance energy requirements (Baldwin et al., 2006; Tovar-Luna et al., 2007). If maintenance energy requirements tend to be lower, then a greater portion of energy intake would be available for productive functions. Another potential explanation for improved efficiency may involve modification of gene expression. It is probable that maternal nutrient supply during gestation, as described by Caton and Hess (2010), may have an influence on the progeny. An example of an occurrence of altered gene expression in a production scale experiment initiated at Fort Keogh Range and Livestock Research Laboratory in the autumn of 2001 (Roberts et al., 2009). At initiation of the study, heifers were developed in the winter with a typical nutritional management scheme or feed 20% less. All heifers were treated the same after the wintering period. Daughters of these cows have also been factorialized across the same 2 winter treatments. Heifers fed the traditional nutritional management scheme have a higher pregnancy rate and greater BW than contemporaries fed 20% less. Other difference have become apparent (Figure 1) after 5 calf crops from daughters who themselves were fed either traditional or conservative and born from dams fed either traditional or conservative. Cows raised by conservatively fed dams and who themselves were conservatively fed have the greatest productivity and longevity while conservativefed cows from conventional-fed dams had lowest longevity. This difference in longevity due to difference in treatment of dam provides evidence that gene expression is altered in utero. It is possible that this changed expression is maintained throughout their life and may potentially be expressed in the daughters’ progeny. Other improvements in efficiency may occur through fine tuning our understanding and management of nutrient and (or) metabolic limitations. Both Lardy and Endecott (2010) and Kerley (2010) described positive changes occurring when postpartum cows were supplied with supplemental MP. Metabolizable protein could be utilized to satisfy MP requirements or go towards supporting gluconeogenesis, as suggested by Waterman and Butler (2010). Kerley (2010) suggested the potential for responses to specific AA, especially methionine. Waterman et al. (2006) reported improved N retention with increasing methionine up to 15 g/d for cows fed typical low-quality roughage diets. Relatively small amounts of AA will lend themselves to smaller, more biologically potent supplements by weight. Smaller supplement should have lower storage, delivery, and feeding costs. A method to determine nutritional adequacy is to monitor and evaluate behavior of grazing livestock. During the conference, we were exposed to improvements in digital technologies that will allow for assessment of precise utilization of which regions of a pasture are used to better model quality of diet intake. These same technologies may be used for a number of grazing behaviors. SUMMARY The speakers and poster presenters provided the participants with an invigorating conference. Future work with the implementation of the alkane technique in the US is sure to be natural outcome from the information received. The ability to link intake measurement with digital tracking and assessment of behavior, especially in the elucidation of grazing distribution, should lead to major advances in the field. There has been a clear progress in our knowledge concerning metabolic dysfunctions relating to lactation and pregnancy. The impact on progeny has better definition. 185 The implication of management on gene expression is a fascinating area that is rich to pursue in the future. The importance of grazing livestock nutrition was easy to realize during the conference. It has an enviable position where it is a field utilizing the most current breakthroughs in science while using those findings for practical applications to improve the lives of producers and consumers and enhances the sustainability of grasslands. biomarkers of longevity, metabolic adaptation, and oxidative stress in overweight individuals - A randomized controlled trial. J. Amer. Med. Assoc. 295:1539-1548. Kerley, M. 2010. Potential for nutritional imbalance in high-quality forages. Pages 125–130 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Lardy, G. P. and R. L. Endecott. Strategic supplementation to correct for nutrient imbalances. Pages 152–167 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Lodge-Ivey, S. L. 2010. Practical application of rumen microbiological techniques to grazing ruminants. Pages 24–28 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Mulliniks, J. T., S. H. Cox, M. E. Kemp, R. L. Endecott, R. C. Waterman, D. M. VanLeeuween, and M. K. Petersen. 2009. Increasing glucogenic precursors in range supplements improves reproductive efficiency and profitability in young postpartum range cows in years 2000-2007. Proc. West. Sect. Amer. Soc. Anim. 60:7680. Roberts, A. J., E. E. Grings, M. D. MacNeil, R. C. Waterman, L. Alexander, and T. W. Geary. 2009 Implications of going against the dogma of feed them to breed them. Proc. West. Sect. Amer. Soc. Anim. Sci. 60:85-88. Torell, L. A. and N. R. Rimbey. Economically efficient supplemental feeding and the impact of nutritional decisions on net ranch returns. Pages 170–177 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Tovar-Luna, I., A. L. Goetsch, R. Puchala, T. Sahlu ,G. E. Carstens, H. C. Freetly, and Z. B. Johnson. 2007. Effects of moderate feed restriction on energy expenditure by 2-year-old crossbred Boer goats. Small Rumin. Res.72:25–32. Waterman, R. C. and W. R. Butler. 2010. Metabolic signals of the beef cow in negative energy balance. Pages 93– 101 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Waterman, R. C., C. A. Löest, W. D. Bryant, and M. K. Petersen. 2006. Supplemental methionine and urea for gestating beef cows consuming low quality forage. J. Anim. Sci. 84:433-446. Waterman, R. C., E. E. Grings, T. W. Geary, A. J. Roberts, L. J. Alexander, and M. D. MacNeil. 2007. Influence of seasonal forage quality on glucose kinetics of young beef cows. J. Anim. Sci. 85:2582-2595. LITERATURE CITED Anderson, D. M. 2010. Geospatial methods and data analysis for assessing distribution of grazing livestock. Pages 57–90 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Baldwin, R. L., VI, K. R. McLeod, and A. V. Capuco. 2004. Visceral tissue growth and proliferation during the bovine lactation cycle. J. Dairy Sci. 87:2977–2986. Brulc, J. M., C. J. Yeoman, K. E. Nelson, and B. A. White. 2010. Emerging methods in rumen microbiology. Pages 10–21 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Caton, J. S. and B. W. Hess. 2010. Maternal plane of nutrition: Impacts on fetal outcomes and postnatal offspring responses. Pages 104–122 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. DelCurto, T. and K. C. Olson. Issues in grazing livestock nutrition. Pages 1–7 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Dove, H. 2010. Keynote: Assessment of intake and diet composition of grazing livestock. Pages 31–54 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Dove, H., S. G. McLennan, and D. P. Poppi. 2010. Application of nutrient requirement schemes to grazing animals. Pages 133–149 in Proc. 4th Grazing Livestock Nutrition Conference. B. W. Hess, T. DelCurto, J. G. P. Bowman, and R. C. Waterman ed. West. Sect. Am. Soc. Anim. Sci., Champaign, IL. Hawkins, D. E., M. K. Petersen, M. G. Thomas, J. E. Sawyer, and R. C. Waterman. 2000. Can beef heifers and young postpartum cows be physiologically and nutritionally manipulated to optimize reproductive efficiency? J. Anim. Sci. 77:1-10. Heilbronn, L. K., L. de Jonge, M. I. Frisard, J. P. DeLany, E. Larson-Meyer, J. Rood, T. Nguyen, C. K. Martin, J. Volaufova, M. M. Most, F. L. Greenway, S. R. Smith, W. A. Deutsch, D. A. Williamson, and E. Ravussin. 2006. Effect of 6-month calorie restriction on 186 Table 1. Four years of yearling heifer development on the ranch (supplemented with one of two protein supplements varying in ruminal degradability) vs. drylot Measurement RUP Development1 RDP Drylot SEM P-value Heifer BW, kg Oct (weaning wt) 224 223 224 3 0.97 Jan/Feb 256 255 255 3 0.98 April/May 276 276 315 4 < 0.01 Sept/Oct (palpation wt) 393 402 403 5 0.24 ADG, kg/d Jan to May 0.26 0.27 0.69 0.02 < 0.01 May to Sept 0.80 0.85 0.61 0.01 < 0.01 Jan to Sept 0.58 0.62 0.64 0.01 < 0.01 Heifer hip height, cm Jan/Feb 116.38 116.64 116.13 0.53 0.80 April/May 120.80 120.88 118.49 0.46 < 0.01 Sept/Oct 123.80 124.76 124.82 0.56 0.32 Heifer BW:HH ratio, kg/cm Jan/Feb 2.20 2.19 2.19 0.02 0.94 April/May 2.28 2.28 2.65 0.01 < 0.01 Sept/Oct 3.16 3.20 3.21 0.02 0.14 Hip height change, cm Jan to May 4.42 4.27 2.34 0.30 < 0.01 May to Sept 2.97 3.89 6.32 0.36 < 0.01 Jan to Sept 7.39 8.13 8.66 0.38 < 0.01 Reproductive performance Pregnancy rate, % 94 88 84 -0.10 Ratio 60/64 56/64 53/63 -1 On-ranch supplements varied in ruminal degradability; RUP = rumen-undegraded protein and RDP = rumen-degraded protein. Table 2. Body weight and body condition responses of gestating cows to different supplementation strategies1 Item CSM SMP NC SE2 P-value Body Weight Responses Weaned BW3, kg 488 492 490 17 0.98 Initial BW4, kg 522 521 534 14 0.78 Final BW, kg 522 523 521 14 0.99 Predicted BW change kg -54.3 -62.8 -64.7 Observed BW change5, kg -0.2a 1.8a -12.6b 3.9 0.06 a a Observed BW change, % 0.1 0.5 -2.2b 0.8 0.09 Body Condition Responses Initial BCS 5.0 4.9 5.0 0.1 0.49 Final BCS 4.9a 4.9a 4.6b 0.1 0.12 BCS change -0.1a -0.1a -0.4b 0.1 0.10 Total feed consumed, kg 38.2 13.7 2.6 --1 Cows received 3.2 kg per week of a 36% CP cottonseed meal-based supplement fed in equal amounts on Monday, Wednesday, and Friday (CSM), 1.6 kg per week of a 65% CP (or higher) animal proteinbased supplement (either blood and feather meal or fish meal) that was self-fed with a mineral mix (SMP), or the CSM at 0.3 kg per week as prescribed at the discretion of the manager depending upon weather conditions (NC). 2 n = 6. 3 Cow BW at weaning in October. 4 Cow BW at initiation of supplementation period. 5 Cow BW change (final BW – initial BW). a,b Means without a common superscript differ (P < 0.10). 187 100% 90% 80% 70% 60% 50% 40% 30% 1 2 C Cdam 3 C Rdam R Cdam 4 5 R Rdam Figure 1. Percent retention of cows over five years from the cow longevity study of Roberts et al. (2009): CCdam = conventional fed daughter and dam; CRdam= conventional fed daughter and restricted dam; RCdam = restricted fed daughter and conventional dam; and RRdam= restricted fed daughter and restricted dam. 188 CHALLENGES TO PREDICTING PRODUCTIVITY OF GRAZING RUMINANTS: WHERE TO NOW? M. K. Petersen, J. T. Mulliniks, A. J. Roberts, and R. C. Waterman Notes Abstracts — Grazing Livestock Nutrition Conference Friday, July 9, 2010 POSTER PRESENTATIONS 1 EFFECTS OF RESTRICTING TIME AT PASTURE ON FEEDING STATION BEHAVIOR OF CATTLE DURING THE FIRST GRAZING SESSION OF THE DAY. P Gregorini*, K McLeod, C Clark, C Glassey, A Romera, and J Jago, DairyNZ, Hamilton, Waikato, New Zealand. Information about short-term behavioral adaptation of cattle to time restrictions at pasture is lacking. The feeding station is the area of pasture an animal can reach at each eating step (ES). This work explored feeding station behavior of cattle during the first grazing session of the day (0800–1200 h) in response to daily restrictions in time at pasture. Six groups of eight Holstein-Friesian cows (470 ± 47 kg, 35 ± 9 days in milk) were strip-grazed on a perennial ryegrass (Lolium perenne L.) pasture for either 4 h after each milking (2 4), 8 h between milkings (1 8), or the whole 24 h period excluding milking times (control). Herbage allowance was 0.07 kg herbage DM/kg live weight/d for all treatments. Mean pre-grazing herbage mass was 3,191 kg DM/ha. The new daily strip of pasture was allocated at 0800 h (after morning milking). Six focal cows per replicate (group) were equipped with Icetags® 3D motion sensors (Ice Robotics, Ltd.) and GPS collars (Data Carter Ltd.) to measure the total distance walked while eating, the number of ES (a.k.a. feeding station) per min (ESR), and calculate the eating step length (ESL, m). Eating behavior was defined as cows with heads down and engaged in acquiring herbage into the mouth. The ESL was calculated by dividing the length of a straight line by the number of steps taken on it. This line was determined by two GPS data points while cows were eating. Bite rate (BR, bites per min) and eating time (ET, min) were determined by six trained observers, randomly assigned to each group and focal cows. Bite rate and ESR were determined at four fixed times during the grazing session for all cows over three consecutive weeks. Eating time was determined every 2 min by scan-sampling, also during the grazing session for all cows over the same three consecutive weeks. The area of feeding stations was calculated as the product of ESL and cows neck length. Bites per ES (BES), velocity of walking while eating (VWE, m per min), total ES (TES), total distance walked while eating (DWE, m), and area explored while eating (AE = area of feeding station TES) were calculated from measurements. Treatment effect was analysed using mixed models , including group of cow as a random effect. Group of cows was the experimental unit. GenStat 11.1 was used for statistical analysis. Results are presented in Table 1. Restricting time at pasture leads to differential food deprivation and consequently different stimuli to eat at the time to face the pasture. The results of this experiment demonstrated that dairy cows reacted to such stimuli mainly by changing locomotion behavior (step length, velocity of walking while eating and distance walked while eating) and eating time, which led to considerable changes in the net area explored while eating (AE). At a constant area allocation per cow (e.g. 100 m2 in the present experiment) and assuming that bite depth is proportional to the sward surface height, the evaluation of feeding stations dynamics permits a general estimation of herbage depletion either in mass or proportion or the sward surface height. This study of foraging behavior contributes to the understanding of the adjustment of eating and moving behavior of cattle to hunger. Feeding station behavior of dairy cows during the first grazing session of the day as affected by restricting time at pasture, either 4 h after each milking (2 × 4), 8 h between milkings (1 × 8) or none (Control). Treatments 18 24 Control SED P value BR (bites/min) 52.1 54.9 46.5 0.54 0.001 ESR (ES/min) 3.3 3.4 2.9 0.32 0.757 BES (bites/eating step) 13.9 14.9 15.0 1.47 0.328 ESL (m) 0.99 0.73 0.82 0.09 0.034 VWE (m/min) 3.3 2.5 2.7 0.26 0.041 ET (min) 195.6 164.4 141.4 5.04 0.004 TES 646.1 544.9 401.2 39.68 0.019 DWE (m) 644.2 389.3 334.3 31.96 0.004 AE (m2) 516.9 311.4 267.4 25.57 0.004 Key words: cattle, pasture restriction, foraging behavior 191 2 SPATIAL AND TEMPORAL FREE-RANGING COW BEHAVIOUR PRE AND POST-WEANING. D.M. Anderson*1, C. Winters1, M. Doniec2, C. Detweiler2, D. Rus2, and B. Nolen1, 1USDA-ARS Jornada Experimental Range, Las Cruces, NM, USA, 2Massachusetts Institute of Technology, Cambridge, MA, USA. Global positioning system (GPS) technology can be used to study free-ranging cow behaviors. GPS equipment was deployed on each of ten cows ranging in age between 3 and 15 years to compare and contrast mean ± standard errors for pre- and post-weaning travel (m·h-1) in two similar ( 433 ha) arid rangeland paddocks. Data were collected at a rate of one position fix per second between 12 March and 8 April 2009 for a total of 8.7 million raw GPS fixes across both paddocks. In addition, all 10 instrumented cows, 5 per paddock, were observed for a total of 20 h and their behaviors were recorded on a minute by minute basis during daylight observations among 13 d. The observational data were used to characterize rate of travel (m·min-1) into walking, foraging and stationary behaviors. Initially the observed data were merged with the GPS data. Next the merged data were examined to classify the GPS data by path speed (m·s-1). The path speed associated with "stationary" was due to the inherent variability within the uncorrected GPS data used in this study. Intervals of 30 s, 60 s, 120 s and 180 s were evaluated to determine the optimum path speed sampling period to discriminate among walking, foraging and stationary behaviors. Regardless of interval chosen, three distinct peaks were observed. Within the ranges evaluated longer sampling intervals (180 s) appeared to show better peak definition especially for those peaks associated with foraging (0.10 m·s-1) and "stationary" (0.03 m·s-1) behaviors. Shorter sampling periods (30 s) produced smother curves having less distinct peaks, probably as a result of more samples. The fastest movement had a peak at 1.00 m·s-1 and corresponded to periods when the cows were observed to be traveling between points on the landscape. A 1.0 m·s-1 rate of travel is roughly 2.2 miles per hour (mph), a speed used to manually gather these cows. Subsequent analyses were performed using 60 s sampling intervals as a compromise for well defined peaks and troughs, smoothness of the curve, and relatively high frequency counts at the high speed end of the distribution. Overall, mean cow travel increased post-weaning over 5 d in Paddock 10B from 1,428 ± 92.6 m·h-1 to 1,955 ± 143.4 m·h-1 and over 8 d post-weaning in Paddock 14A from 1,166 ± 105.8 m·h-1 to 1,509 ± 92.0 m·h-1. Mean travel in the two paddocks during foraging both pre- and post-weaning varied within a day but not in an identical manner even though the two paddocks were similar in size and topography. Among four 6 h time intervals foraging travel decreased following weaning between 6 am and noon and between 6 pm and midnight in both paddocks while travel associated with foraging increased between noon and 6 pm in both padocks. Foraging travel between midnight and 6 am decreased in Paddock 10 following weaning (64 ± 5.6 m·h-1 vs 42 ± 4.4 m·h-1) but increased in Paddock 14A (24 ± 6.6 m·h-1 vs 60 ± 7.3 m·h-1) during this same time interval following weaning. Differences in foraging travel between midnight and 6 am in the two paddocks may have resulted from several factors including: the number of non-instrumented cows being different in the two paddocks, different weaning dates 13 d apart, and a different spatial distribution of cows within each paddock. Furthermore, GPS data were lost possibly due to battery failure. Only one 24 h period of data were lost from one cow in Paddock 10 B (post-weaning) compared to loss of GPS data from two and three of the five cows for 24 h pre- and post-weaning, respectively, in Paddock 14A. Further data analysis evaluating tortuosity of travel (a metric that relates path speed or the summation of distances between each of 60 positions within a minute to displacement speed the distance between only the first position and the 60th position within a minute) appears promising to further characterize free-ranging cow travel. Key words: range animal ecology, beef cow travel, GPS tracking 192 3 EFFECT OF SUPPLEMENT ENERGY LEVEL ON FATTY ACID PROFILES AND MEAT QUALITY OF STEERS FINISHED ON WINTER ANNUAL PASTURE. Harold Ospina Patino* and Fabio Schuler Medeiros, Universidade Federal de Rio Grande do Sul, Porto Alegre, RS, Brazil. Twenty four Aberdeen Angus Charolais steers were used in a completely randomized design to evaluate the effect of supplemental energy levels on performance, meat quality and fatty acid profile of intramuscular fat. The experimental treatments were levels of feeding (0, 0.4, 0.8 and 1.2 kg DM/100 kg BW/day) of a corn based supplement in a winter pasture of annual ryegrass (Lolium multiflorum L.) and oats (Avena strigosa Schreb) managed for warrant higher forage allowance (15 kg DM/100 kg BW/day). Animals were supplemented daily (14:00 – 16:00 h) in individual pens and slaughtered when they achieved 4.5 mm of fat cover on the rump point (ultrasound evaluation). No differences were observed in fat deposition, measured in live animals with ultrasound at the end of the performance period, and in live weight gain, which had average values of 3.9 mm and 1.54 kg/day, respectively (P>.05). Increasing levels of energy supplementation did not influence tenderness (3.36 kg/cm2, marbling (5.63 points), pH (5.64), lipid concentration (9.38%), Color a (15.25), Color b (12.84) in the Longissimus dorsi samples (P>.05). Sample lightness (L) were linearly increased with increasing levels of supplementation on pasture (P=.05). Supplementation level linearly decreased n-3 FA ( P<.001) and linearly increased n-6:n-3 ratio(P=.0011). Concentration of CLA decreased with increased supplementation level (P<.001). Fattening animals on winter pastures using increasing levels of energy supplementation does not result in differences in meat quality but results in changes in fatty acids profile of intramuscular fat. Key words: winter pasture, fatty acid profile, energy supplementation 193 4 USE OF N-ALKANES AND LONG CHAIN ALCOHOLS TO ESTIMATE FORAGE INTAKE AND DIET COMPOSITION OF CATTLE GRAZING TWO FORAGE SPECIES. H.T. Boland*1 and G. Scaglia2, 1 Prairie Research Unit, Mississippi State University, Prairie, MS, USA, 2Iberia Research Station, Louisiana State University, Agricultural Center, Jeanerette, LA, USA. Estimation of dry matter intake (DMI) and diet composition of livestock is a valuable area of study due to its impact on the health, nutritional status, and productivity of animals. It also helps to increase our knowledge of how foraging behavior can impact biodiversity of plant communities. The use of long chain saturated hydrocarbons (n-alkanes) has been shown to be a non-invasive and accurate method to estimate DMI of grazing livestock. In recent years, other components of the cuticular wax of plants, such as long chain alcohols (LCOH) have been explored for use in the estimation of diet composition of animals grazing mixed diets. This technique is possible because the patterns of n-alkane and LCOH concentrations in forage species differ, particularly between grasses and legumes. When animals must search for a preferred forage within a mixed sward, the animal's ability to eat what it wants when it wants becomes constrained. Within this constraint, diet preference of an animal cannot be evaluated. However, this constraint can be minimized when forages are offered as monocultures in adjacent areas. The objectives of the present study were: 1) to determine the DMI of beef steers grazing adjacent monocultures of tall fescue [Lolium arundinaceum (Schreb.) Darbysh. = Festuca arundinacea Schreb.] and alfalfa (Medicago sativa L. ssp. sativa) across two years using naturally occurring and dosed n-alkanes, and 2) to determine diet composition using nalkanes and LCOH. Twenty-four Angus-crossbred steers (12 steers yr-1, Yr 1 initial BW = 392 ± 8 kg, Yr 2 initial BW = 323 ± 9 kg; 16 months old) were used. Each steer was dosed with an intra-ruminal n-alkane (C32 and C36) controlled release fecal marker capsule. After 7 days, forage and fecal samples were collected twice daily for the next 7 days. Purified n-alkane and LCOH extracts from those samples were quantified by gas chromatography. Diet composition was then estimated using a non-negative least squares procedure ('EatWhat' software program). Data were also analyzed using SAS Proc MIXED (ver. 9.1; Cary, NC). The model included yr, forage type, and their interaction for analysis of marker profiles and nutritive value. The model inclued yr, day, and their interaction for the analysis of diet composition and DMI. The repeated measure was day within yr. The n-alkane profiles of tall fescue and alfalfa were similar to those previously reported and there were effects of yr (P 0.07) for the majority of chain lengths within forage species. Higher concentrations of n-alkanes in Yr 2 may have been due to high air temperatures and drought conditions in that year. The LCOH profile of tall fescue was also similar to previously reported data; however information on the LCOH profile of alfalfa had not been reported at the time of this analysis. There were effects of yr (P 0.09) on LCOH profile of forages however the pattern of effect (whether an increase or decrease in concentration) varied between forage species and chain lengths. The use of LCOH added additional characterization of the forages, but in this case diet composition estimates were not different (P 0.22) than when estimated using four different n-alkanes. Diet composition estimation indicated that steers consumed similar (P = 0.13) diets of 79 and 70% alfalfa in Yr 1 and Yr 2, respectively. Dry matter intake differed (P = 0.002) between years with values being lower (4.5 kg d-1, 1.4% BW) in Yr 2 than Yr 1 (9.4 kg d-1, 2.3% BW) which may be explained by differences in weather conditions. Results suggest that if n-alkane profiles of the forages being grazed are different, the additional analysis needed to determine LCOH concentrations may not be necessary. However, in pastures with greater plant species diversity, the use of LCOH or other additional plant-wax markers would most likely be required to estimate diet composition. Key words: n-alkanes, long chain alcohols, diet composition 194 5 PROTEIN AND CARBOHYDRATE DEGRADATION CHARACTERISTICS AND RATIOS OF ANTHOCYANIDIN-ACCUMULATING LC-ALFALFA AND ALFALFA SELECTED FOR A LOW INITIAL RATE OF DEGRADATION IN GRAZING CATTLE. A Jonker*1,2, M Gruber2, Y Wang3, and P Yu1, 1 Department of Animal and Poultry Science, University of Saskatchewan, Saskatoon, SK, Canada, 2Saskatoon Research Centre, Agriculture and Agri-Food Canada, Saskatoon, SK, Canada, 3Lethbridge Research Centre, Agriculture and Agri-Food Canada, Lethbridge, AB, Canada. Alfalfa suffers from the disadvantage of having an excessive rapid protein (CP) degradation which leads to poor protein efficiency and bloat in grazing cattle. Accumulation of monomeric and polymeric anthocyanidins in alfalfa forage or increased alfalfa plant cell wall content have been found to reduce protein degradation in the rumen. The objective of this study was to determine CP and carbohydrates (CHO) degradation and synchronization in the rumen of newly developed anthocyanidin-accumulating T1Lc-alfalfa populations and an alfalfa variety called AC Grazeland that was selected for a low initial rate of degradation (LIRD). Three winter hardy alfalfa varieties Beaver, Rambler and Rangelander were crossed with three T0 Lc-alfalfa genotypes to generate progenies T1BeavLc1, T1RambLc3 and T1RangLc4, respectively. Purple T1Lc-progeny from each of these three populations were compared with LIRD-alfalfa, which was previously found to reduce the incidence of bloat in grazing cattle. Alfalfa samples were collected at a vegetative pre-bud stage during the growing season of 2008 (AAFC, Saskatoon, SK, Canada). An in situ trial was performed in duplicate runs and results inserted into an exponential model to determine ruminal degradation characteristics and synchronization. The results were analyzed in a completely randomized design using Proc Mixed of SAS. The T1Lc-alfalfa populations had an average anthocyanidin concentration of 163.4 µg/g DM while LIRD-alfalfa did not accumulate anthocyanidin. Undegradable CHO and CHO residues at 36 and 72 h of ruminal incubation were lower (P<0.05) in T1Lc-alfalfa compared with LIRD-alfalfa. Protein residue at 36 and 72 h of ruminal incubation was lower (P<0.05) in T1Lc-alfalfa compared with LIRD-alfalfa. Total rumen degradable protein tended to be higher (P<0.10) in T1Lc-alfalfa compared with LIRD-alfalfa (225.6 vs. 213.8 g/kg DM). Total rumen degradable N to CHO ratio was lower (P<0.05) in T1RambLc3 compared with T1BeavLc1 which had and lower (P<0.05) N to CHO ratio compared with T1RangLc4 (110.2 vs. 115.2 vs. 118.8 g/kg) while LIRDalfalfa was not (P>0.05) different from T1BeavLc1 and T1RambLc3. Hourly rumen degradable N to CHO ratio was similar (P>0.05) between the four populations with mean values of 197.9, 110.3, 87.3 and 59.9 g/kg after 1, 2, 6 and 12 h of incubation, respectively. In conclusion, T1Lc-alfalfa accumulated anthocyanidin and had similar degradation profiles and ratios compared to LIRD-alfalfa which was earlier found to be bloat reduced in grazing cattle. Thus, a further increase in monomeric and polymeric anthocyanidins accumulation in alfalfa is required in order to develop an alfalfa variety with superior nutritional and bloat preventing characteristics compared to currently available varieties. Key words: anthocyanidin-accumulating alfalfa, in situ ruminal degradation characteristics, grazing cattle 195 6 GROWING AND FATTENING CATTLE ON THE NORTHERN GREAT PLAINS WITH RUMINALLY-PROTECTED FLAXSEED. S. L. Kronberg*1, E. J. Murphy2, R. J. Maddock3, and E. J. Scholljegerdes1, 1USDA-ARS, Northern Great Plains Research Laboratory, Mandan, North Dakota, USA, 2 Departments of Pharmacology, Physiology and Therapeutics, and Chemistry, University of North Dakota, Grand Forks, North Dakota, USA, 3Department of Animal Sciences, North Dakota State University, Fargo, North Dakota, USA. Feeding ruminally-protected flaxseed may help minimize time in a feedlot and improve profitability and sustainability of beef production and healthfulness of beef. Twelve yearling steers grazed together on prairie grassland from June to early August then on August 16th were weighed (452.8 ± 15.5 kg) and randomly divided into two groups: l) received ground flaxseed (908 g/steer/d; FLAX; n = 6); 2) received protected flaxseed at the same rate (PFLAX; n = 6). Steers were individually fed the treatments. Steers grazed together on Proso millet from August 16th to September 26th, then on irrigated smooth brome-dominated pasture until October 10th. From October 11th to November 14th they were penned and each morning individually fed their treatments along with rolled barley (6.36 kg/steer/d for final 27 d of trial) then given free-choice access to prairie grass hay (all steers together) for remainder of day. Steers were weighed at end of the grazing period and at end of the penned period. Growth rates and carcasses were evaluated. Samples of LM muscle were collected, muscle fat was separated into neutral and polar lipids, and fatty acid (FA) profiles determined with GLC. There was no difference between the treatment groups for ADG while grazing (P = 0.86; 1.38 kg/d), nor while penned (P = 0.46; 1.76 kg/d). No differences between groups were observed for HCW (P = 0.37; 322.8 kg), USDA yield grade (P = 0.25; 2.4), fat thickness over 12th rib (P = 0.77; 7.4 mm), or marbling score (P = 0.42; 394). However, 67% of PFLAX steers graded USDA High Select or greater quality grade; whereas, only 33% of FLAX steers did this. No differences (P 0.41) were observed between groups for levels (mole %) of 18:3n-3, 20:5n-3, 22:5n-3 or total n-3 FA in the polar lipid fraction of LM muscle, but levels of 16:0 and total saturated FA were higher (P 0.02) in the neutral lipid fraction of LM muscle of FLAX, while levels of total unsaturated FA and monounsaturated FA were higher (P 0.01) in the neutral lipid fraction of LM muscle of PFLAX. Feeding treated flaxseed to yearling steers did not increase their growth rate or levels of n-3 FA in LM muscle, but appeared to improve USDA quality grade and decreased levels of total saturated FA while increasing levels of total unsaturated FA and monounsaturated FA in the neutral lipid fraction of LM muscle. Key words: fatty acids, Omega-3 fatty acids, beef 196 7 GRASS INTAKE OF GRAZING DAIRY COWS USING THE N-ALKANE TECHNIQUE. A. van den Pol-van Dasselaar*1 and A. Hensen2, 1Wageningen UR Livestock Research, Lelystad, the Netherlands, 2Energy research Centre of the Netherlands, Petten, the Netherlands. Grass intake of grazing dairy cows can be estimated using the n-alkane technique. N-alkanes are saturated, aliphatic hydrocarbons with lengths varying from 21 to 37 carbon atoms. They are part of the cuticular wax of plant leaves. When synthetic alkanes are added to the diet in a fixed dose, the herbage intake in the field can be estimated using the ratio between natural and synthetic alkanes in the faeces of grazing animals. Two experiments were carried out. A first experiment was carried out to investigate the effectiveness of the n-alkane technique for rations of grass, maize silage and concentrates. The n-alkane technique has shown to deliver reliable estimates of dry matter (DM) intake during grazing, but no information was available on the recovery of diets of grass supplemented with maize silage. In the Netherlands, rations of grass and maize silage are common. A second experiment was carried out to study the relationship between grass intake, as estimated by the n-alkane technique, and methane emissions of dairy cows in a grazing situation with supplemental feeding. The first experiment was carried out indoors with four productive dairy cows. Two rations were compared: a ration of fresh grass and concentrates and a ration of fresh grass, maize silage and concentrates. The concentrates were enriched with the synthetic alkanes C32 and C36. Feeds and manure were analysed as to the relevant alkanes. Maize silage showed much lower n-alkane levels than grass (3 to 15 times lower). Individual alkane recoveries of C25, C27, C29, C31, C32, C33, C35 and C36 were estimated by total faecal collection. The results showed that the ration may affect recoveries of n-alkanes. The recoveries of the shorter alkanes (especially C25 and C27) decreased significantly by supplementation of maize silage (P<0.1). Recoveries of alkanes like C32 and C33 were not significantly affected by the inclusion of maize silage in the ration. For a reliable application of the n-alkane technique, the recovery of the individual alkanes is crucial. Using the shorter alkanes will not give reliable estimates for intake of rations of grass and maize silage. The second experiment was carried out in the field. We used C32 and C33 to measure herbage intake. There were three measurement periods in two consecutive years. In the first year of the experiment, five productive dairy cows were used. In the second year, six productive dairy cows were used. The animals stayed indoors during night and grazed during the day. Their milk production was on average 25 kg milk per day. Indoors they were fed a ration of 5 kg DM maize silage per day and 3 kg DM concentrates per day. The concentrates were enriched with a C32 alkane marker. Grass intake during grazing was estimated using the n-alkane technique using the daily dose of C32 alkanes from the concentrates, the concentration of natural occurring C33 alkanes in the grazed grass and the ration C32:C33 in the faeces. Furthermore, methane emissions of the grazing dairy cows were measured using the sulphur hexafluoride (SF6) tracer technique. Grass intake per animal varied between 8 and 13 kg DM per day. Grass intake was positively related to methane emissions. The base-line of methane emissions differed between the measurement periods. On average, however, per kg additional DM intake an extra 35 g CH4 per animal per day was emitted (r2=0.6-0.7). The animals with a high DM intake generally also produced more milk. This led to less methane emitted per kg milk for animals with a high DM intake compared to animals with a low DM intake. Other research confirms that high productive dairy cows emit less methane per kg milk than low productive dairy cows. We conclude that using the n-alkane technique with a combination of C32 and C33 is suitable to provide reliable estimates of fresh grass intake in situations with supplemental feeding. Using this n-alkane technique, we found a positive response between DM intake and methane emissions of grazing dairy cows. Key words: grass intake, n-alkane technique, grazing 197 8 GRAZING IN EUROPE. A. van den Pol-van Dasselaar*, Wageningen UR Livestock Research, Lelystad, the Netherland. Throughout Europe, forage is the main feed for dairy cattle. Grass is fed either fresh - predominantly through grazing - or in a preserved form as silage or hay. Long-term data of grazing in Europe are hardly available. The objective of this study was to provide insight in the state-of-the-art with respect to grazing in Europe. In 2010, a survey among scientists of several countries in Europe was carried out. Leading scientists related to the Working Group Grazing of the European Grassland Federation were asked to provide data on grazing. In most countries, statistical data on this subject were not available. In those cases, the scientists were asked for their expert judgement. The survey indicated that zero-grazing becomes more and more popular. There is a large variation in percentage of grazing dairy cows, both between and within countries, e.g.: Finland, 80% grazing, slow decrease, Sweden, 100% grazing, defined by law, Denmark, 35-45% grazing, strong decrease, the Netherlands grazing, 79%, decrease, Belgium, 90% grazing, decrease, Luxembourg, 85% grazing, strong decrease, NW Spain, 15% grazing, slow increase, Slovenia, 25% grazing, stable or lower, Greece, 15% grazing, slow increase. Grazing systems used differ between countries. The survey revealed that rotational grazing is practised the most often. When grazing is practised, cows graze mainly during the day. During the night, cows are indoors and get supplemental feeding. The number of hours grazed per year and per day is decreasing. In some countries (e.g. the Netherlands, Denmark, Belgium) a new development in grazing is the mobile automatic milking system, which is able to milk the animals in the field. This system is currently being tested in research projects. In Northern Europe, grazing is practised more often than in Southern Europe. However, also in Northern Europe the percentage grazing is decreasing rapidly. Is this a matter of concern? The grazing system affects various aspects like grassland productivity, animal welfare, environment, economy, labour and even society. Grazing has advantages and disadvantages (see Table). The importance attached to the various effects of grazing is very personal. Restricted grazing scores well on the whole. We expect the trend of less grazing in Europe to continue. There are economical, practical and personal motives behind this. In the end, the personal preferences of the farmer determine the grazing system used. The decline in the popularity of grazing is supported by current trends in livestock farming in Europe. Average herd sizes increased during the last years and the number of automatic milking systems increased. Grazing of large herds is difficult to manage. And even though grazing in combination with an automatic milking system is possible, it is experienced as difficult. The average milk production per cow increased and farmers with high yielding cattle like to control rations. Again, control of rations is more difficult in grazing situations. In countries where grass growth is delayed in summer, another reason for less grazing may be the uncertainty whether grass growth suffices for the demand of grazing cows. Finally, reasons for zero-grazing may be better grassland utilisation, the need to reduce mineral losses and ease of labour. Especially the latter is an important factor for many farmers. In most countries, there are hardly any tools available to support farmers in grazing. In conclusion, a survey among scientists in Europe showed that the number of grazing dairy cows is rapidly decreasing. Simple and easy-to-use grazing systems and practical management tools have to be developed to support farmers in grazing. The effect of grazing (unrestricted grazing, restricted grazing, zero-grazing) on various aspects. The score ranges from - to ++, with ++ signifying that the system concerned scores positive for the point in question, e.g. high health, low losses. Unrestricted Restricted Zero Grass yield and grass use - + + Balanced diet - +/- ++ Natural behaviour ++ ++ + Animal health ++ + +/- Nitrate leaching, N2O emission - + ++ N losses - + ++ P losses - +/- + Ammonia volatilisation ++ + +/- Energy use, CH4 emission + - -- Fatty acid composition of milk ++ + +/- Labour/hours work per year ++ + + Economics + + - Image of dairy farming ++ + - Key words: grazing, Europe, grazing systems 198 9 PATHWAY FOR THE ELIMINATION OF MELAMINE IN LACTATING DAIRY COWS. Junshi Shen, Jiaqi Wang*, Hongyang Wei, Dengpan Bu, and Peng Sun, State Key Laboratory of Animal Nutrition, Institute of Animal Science, Chinese Academy of Agricultural Sciences, Institute of Animal Science, Chinese Academy of Agricultural Sciences, Beijing, China. Melamine (MEL) might be degraded into cyanuric acid (CYA) and some other analogues by the rumen microorganism. The objective of this study was to investigate the pathway for the elimination of MEL in dairy cows. Four late-lactating dairy cows (Body weight = 564 ± 17 kg, Days in milk = 265 ± 14d, Milk yield = 13.2 ± 2.16kg) fitted with ruminal cannulas were dosed with MEL (purity 99.5%) at 800 mg/d per cow, divided into two equal daily doses. The trial lasted for 20 d. The first 13 d was preliminary period, followed by a 7-d sample collecting period, during which feed intake and total output of feces, urine, and milk were measured from d 14 to 16. Blood samples were collected at 0 (pre-dose), 4, and 8 h after morning feeding on d 17 and 18, and rumen fluid were collected at 0 (pre-dose) 1, 2, 4, and 8 h after morning feeding on d 19 and 20. Plasma MEL or CYA concentration was analyzed using the GLM procedures of SAS system. The percentage of MEL excreted through milk, urine, or feces was calculated as the ratio between the mean output of MEL in milk, urine, or feces and the MEL administered per day, separately for each cow. Method of Liquid chromatography and tandem mass spectrometry (LC-MS/MS) was utilized to determine MEL and CYA simultaneously. Before the trial started, no MEL or CYA was detected in samples of feed, milk, plasma, urine and feces. MEL concentration in rumen fluid (Y, mg/L) decreased exponentially after the morning feeding (X, h) (Y = 3.85591e -X/9.25674 + 1.35924, R 2 = 0.99), but no CYA was detected. Plasma MEL concentration was relatively stable at the three different sampling times (P > 0.05). The percentage of MEL excreted through milk, urine and feces were 0.48 ± 0.06, 10.98 ± 3.88 and 44.07 ± 10.79%, respectively. Therefore, 44.47 ± 7.98% of ingested MEL was degraded in the rumen. This speculation was conformed by the fact that CYA was also detected in plasma, urine and feces. Whether ammelide and ammeline were excreted commonly with MEL needs further study. The results of the present work implied that the metabolism of MEL in dairy cows was different from monogastric animals. High percentage of MEL may be degraded gradually by rumen microorganism to produce its analogues (CYA). The digestive tract and kidney are two main pathways for the excretion of MEL through feces and urine. Key words: melamine, cyanuric acid, dairy cow 199 10 COMPARE THE DIGESTIBILITY OF RUMEN UNDEGRADABLE PROTEIN OF CORN DDGS AND DDG IN INTESTINE OF CHINESE HOLSTEIN DAIRY COWS USING MOBILE NYLON BAG TECHNIQUE. Z.H. Yan, J.Q. Wang*, D.P. Bu, H.Y. Wei, L.Y. Zhou, and P. Sun, State Key Laboratory of Animal Nutrition, Institute of Animal Science, Chinese Academy of Agriculture Sciences., Beijing, China. As the bioethanol industry continues to expand, Corn dried distillers grains with solubles (DDGS) and corn dried distillers grains (DDG) become increasing available as animal feed. Three DDGS samples and three DDG samples were collected from ethanol plants in Middle Eastern China. Dry matter and CP disappearance in rumen and small intestine were measured in four Chinese dairy cows (638±61kg) in dry period using a mobile nylon bag technique. Samples of each DDG and DDGS were incubated for 16 h in the rumen; the residues from these bags were transferred to mobile bags. All bags were placed in a 0.1 N HCl solution containing pepsin (1 g/L; Sigma P7000; Sigma-Aldrich, St Louis, MO, USA), and incubated for 1 h at 39°C to mimic abomasal digestion, and then used for determination of intestinal digestibility in situ. Ruminal degradability and intestine digestibility of DM and CP were analyzed using a randomized complete block design with the MIXED procedure of SAS (SAS Institute, 2001). Ruminal degradability of DM and CP for DDGS and DDG did not differed after 16 h ruminal incubation (DM, 42.2 and 49.9% for DDG and DDGS, P=0.12; CP, 31.0 and 33.6% for DDGS and DDG, P=0.59). The ruminal degradation of DM and CP for DDG or DDGS differed between plants (P <0.0001). Ruminal degradation of DM and CP were greater for DDGS1 and DDGS3 compared with DDGS2. Intestine digestibility of DM was greater for DDGS compared with DDG (P=0.001). The intestine digestibility of CP was almost the same for DDG and DDGS (P =0.99). And intestine digestibility of DM and CP for DDG or DDGS differed significantly between plants. Result of this study showed that ruminal degradation differed larger for DDG samples, whereas intestinal digestibility changed more for DDGS. In concern with the nutritional characteristic of DDGS or DDG differed largely in various plants, it would be effective to develop a quick and easy test technique to determine the feed value of the DDGS or DDG that could be used by the farmer. Table 1 Ingredient and nutrient composition of the diets fed to cows during the in situ experiment Ingredients % DM Nutrient composition Alfalfa hay 16 NEL 1 (Mcal/kg DM) 1.57 Chinese wildrye 14 CP, % 17.89 Corn silage 20 NDF, % 36.77 Corn ground 23 ADF, % 22.13 Wheat bran 6 EE, % 3.1 Soybean meal 10.5 Ca, % 0.84 Cotton seed meal 4 P, % 0.48 Rapeseed meal 4 Calcium carbonate 1.2 Calcium phosphate, dibasic 0.3 NaCl Premix 0.5 2 0.5 1 NEL was estimated from the feeding standards of dairy cattle, China NongYe HangYe Biaozhun/Tuijian-34 (China NY/T34, 2004). 2 Premix contained vitamin A >700,000 UI/kg, vitamin D3 >120,000UI, vitamin E>2100mg, Fe 1750mg/kg, Cu 1600mg/kg, Zn 10000mg/kg, Mn 3500mg/kg, Se 42mg/kg, I 84mg/kg, and Co 42mg/kg Key words: DDGS(DDG), ruminal degradation, intestinal digestibility 200 11 EFFECT OF SITE AND SOURCE OF LYSINE SUPPLEMENTATION ON NITROGEN METABOLISM IN BEEF CATTLE. Y.Q. Guo, Y.D. Zhang, J.Q. Wang*, D.P. Bu, K.L. Liu, and T. Hu, State Key Laboratory of Animal Nutrition, Institute of Animal Science, Chinese Academy of Agriculture Sciences., Beijing, China. Four Limousin-cross cows (BW=582±42 kg) fitted with both duodenal and ruminant cannulas were used to investigate the effect of ruminal supplementation of Lysine (15 g/d, RL) or rumen protected Lysine (26 g/d, protected by polyoxylate IV and hydrogenated vegetable tallow containing of Lysine 15g, RPL) and duodenum infusion of Lysine (15 g/d, DL) on N utilization and plasma amino acids in a 44 Latin Square design with each period lasting 7 d. The group of basal diets served as the control. Blood samples were collected on d 7, and ruminal fluid and feces, urine were collected for 3 d from d 4 to d 6 per 7 d period. All statistical analyses were performed using PROC GLM of SAS. The results showed that supplementation with Lysine did not affect N intake (P>0.05), regardless of supplementation site or source. Higher ruminal NH3-N concentration (6.54 mg/dl vs. 5.19 mg/dl, P=0.11), plasma urea N concentration (7.16 mg/l vs. 5.36 mg/l, P=0.02), urinary N excretion (6.51 g/d vs. 6.36 g/d, P=0.64), N retention efficiency (33.41 vs 23.66, P=0.42, g/d as well as percentage of absorbed N) and less fecal N excretion (38.34 g/d vs. 42.40 g/d, P=0.18) were observed when compared with the control. RPL and DL treatments increased plasma Ala, Pro, Tyr, Leu, Phe, Arg and essential amino acids concentrations (P<0.05). The concentrations of Lysine were improved in RPL (11.18 ug/ml) and DL (16.97 ug/ml) vs. RL (8.00 ug/ml) and control (7.31 ug/ml, P=0.0001). The ratio of RPL Lysine concentration to DL Lysine concentration was 66.1% in plasma. The results suggest that supplementation of Lysine could improve N utilization efficiency to some extent and rumen addition of protected Lysine is an effective way to supply Lysine for Limousin-cross cows. Key words: lysine, nitrogen metabolism, beef cattle 201 12 SUPPLEMENTATION WITH PROTEIN AND VARIOUS FEED ADDITIVES IMPROVES IN SITU DEGRADATION CHARACTERISTICS OF BERMUDAGRASS AND BUFFELGRASS IN STEERS CONSUMING A LOW-QUALITY FORAGE DIET. K.C. McCuistion*1, B.D. Lambert2, T.A. Wickersham3, R.O. Dittmar3, L. Wiley1, and L. Dobson3, 1Texas A&M University-Kingsville, Kingsville, TX, U.S., 2Tarleton State University and Texas AgriLife Research, Stephenville, TX, U.S., 3Texas A&M University, College Station, TX, U.S. Ruminant animals face periods of nutrient restriction due to changes in forage quality and quantity often the result of environmental factors including drought. If nutrient inputs through forage quality and intake do not meet animal requirements, optimal production will not be achieved. Supplementation is a risk management tool that can be used to overcome nutrient deficiencies. Supplementation also provides a vehicle for delivery of compounds which alter metabolism and/or improve digestion. The objective of this study was to determine the rate and extent of in situ disappearance of two forages, Bermudagrass (Cynodon dactylon (L.) Pers.) and Buffelgrass (Pennisetum ciliare (L.) Link), when incubated ruminally in steers fed a basal diet of low-quality hay. Ten ruminally cannulated Angus steers (311 ± 28 kg) were used in an incomplete Latin square design. Each steer was randomly assigned to one of five treatments with two steers per treatment and two treatment periods (n=4). Treatments were: 1) control (C; no supplement); 2) protein supplement (CSM; cottonseed meal at 2g/kg BW); 3) protein supplement (2g/kg BW CSM) plus calcium propionate (200 g per d; P); 4) protein supplement (2g/kg BW CSM) and sodium formate (2.5 g per d; SF); and 5) protein supplement (2g/kg BW CSM) and monensin (200 mg per d; M). Supplements were provided once daily to individual steers. In order to simulate nutritionally stressed conditions, low-quality old world bluestem hay (2.7% CP, 82.4% NDF, and 68.0% ADF) was ground to pass through a 75mm 75mm screen and provided ad libitum. Steers had free access to water and a trace mineral salt. Each period lasted 14 d which consisted of a 10 d adaptation to treatment followed by the in situ incubation. Both Bermudagrass (12.8 %CP, 66.5% NDF, and 45.0% ADF) and Buffelgrass (5.7% CP, 73.8% NDF, and 60.3% ADF) were incubated in situ for 0, 2, 4, 8, 16, 24, 48, and 72 h to determine rate and extent of DM disappearance. There was no interaction between treatment and forage variety (P 0.24). In situ disappearance components included the following fractions: a (soluble fraction), b (potentially degradable insoluble fraction), k (hourly fractional disappearance of b) and L (discrete time lag before disappearance of b began). Bermudagrass had a larger a fraction and a faster rate of disappearance for b than Buffelgrass (P < 0.01; Table 1), as would be expected given its higher nutritive value. When total DM disappearance was calculated for each treatment, the C had the lowest amount of a fraction whereas SF had the highest, with CSM, P, and M being intermediate (P = 0.01). Fractional rate of disappearance followed the same trend (P < 0.01), indicating that all supplements and feed additives had a greater rate of forage disappearance than the C. Results indicate that protein supplementation, in addition to certain feed additives, can improve DM disappearance of lowand moderate-quality forage when provided to cattle consuming a basal diet of low-quality forage. Table 1. Rumen in situ DM disappearance characteristics for forage variety and treatment. a (%) b (%) k (/h) L (h) Bermudagrass 34.74x 35.74x 0.033x 3.67x 23.58y 51.17y 0.024y 7.78y 18.31a 54.38b 0.013a 7.40 0.028 b 4.06 0.036 b,c 7.36 Forage Buffelgrass Treatment C 26.03 a,b P 35.23 b,c M 30.14b,c 41.98a CSM SF 36.06 c 46.64 a,b 37.54 a 36.74 a 0.027b 5.89 c 3.91 0.038 1 C = Control; CSM = cottonseed meal protein supplement; P = calcium propionate + CSM; M = monensin + CSM; SF = sodium formate + CSM. x,yMeans not bearing a common superscript letter differ by (P < 0.05) for forage variety. a,b,cMeans not bearing a common superscript letter differ by (P < 0.05) for treatment. Key words: feed additives, low-quality forage, in situ 202 13 CASE STUDY: MOLASSES AS THE PRIMARY ENERGY SUPPLEMENT ON AN ORGANIC GRAZING DAIRY FARM. K. Hoffman1, L.E. Chase2, and K.J. Soder*3, 1USDA-NRCS, Norwich, NY USA, 2 Cornell University, Ithaca, NY USA, 3USDA-ARS, University Park, PA USA. Due to increasing organic grain costs, organic dairy farmers are looking for less expensive ingredients. Molasses seems to be a less expensive source of supplemental energy. Organic dairy farmers inquire about molasses as an alternative based on farmer testimonials. Research has been conducted on molasses as feed for many decades, much of it published in the 1950's to 1970's. Today's knowledge of dairy nutrition is much greater, so results of that research may not be directly applicable. Also, little research was conducted with dairy cows fed a high rate of molasses as the sole energy supplement, or with dairy cows grazing cool-season pastures in the northeastern USA, and limited, at best, research is expected in the next few years, but farmers are seeking answers now. Anecdotal results have been mixed, with some farms reporting success, while others reporting failure. The objective of this case study was to quantify nutrient intake, milk production and animal performance on an organic dairy farm in central New York that is currently feeding molasses. Forage quality, nutritional data, and other management techniques (i.e. pasture management) were collected monthly during monthly farm visits during the 2008 grazing season. Data was transferred to Microsoft Excel files and summarized. Pasture and supplements samples were collected during each farm visit and analyzed for protein, fiber, energy, and mineral content. The Cornell Net Carbohydrate and Protein System (CNCPS) model was used to analyze the diet on a monthly basis. The CNCPS has the ability to predict milk production based upon descriptive parameters of the animals, environment, activity level, and the quality of the diet. In this study, it was used to compare actual performance to predicted performance. Body condition score (BCS) averaged slightly over 2.5 in May (40-60 days in milk). At the lowest point in August, BCS averaged 2.1 and increased to 2.71 by November. Peak milk production occurred in May at 24 kg per cow, and then gradually declined through the rest of the study period. Production in November was 11 kg per cow, which is a persistency rate decline over the 6 month period of 12% per month. However, the rate of decrease in persistency from late May to early July was greater at 25% per month, likely due to the lower pasture quality during this time. Also, breeding began in early June, which may have increased their energy requirements slightly. Pregnancy rate was 95% using natural service, indicating that the cows were in an adequate energy status. Milk components were very consistent throughout the grazing season, ranging from 4.01% to 5.53%. Milk protein and other solids were also consistent, with milk protein averaging 3.4% and trending up towards the end of the study as milk production declined. Further, by November the cows were being fed a fair amount of dry hay and baleage, which would help to increase the milk fat percentage. The milk urea nitrogen levels were relatively high (16.2 mg/dl average) throughout the grazing season, as would be expected on a high protein pasture diet with low levels of non-fiber carbohydrates. Compared with simulated results from the CNCPS, the cows were not as efficient at utilizing the pasture protein as the model expected. These results raise these questions: 1) does molasses provides a higher level of energy compared to corn?; and 2) are the model values for molasses used to predict rumen dynamics accurate? The cost of organic starch sources in comparison to the cost of organic molasses needs to be evaluated both individually and in combinations of various feeding rates, as interest in the use of molasses is motivated by higher organic grain costs. This study was continued through the 2009 grazing season, as the farm decided to feed a slightly higher grain rate due to some cows beginning their lactation in low body condition, the potential for better milk production, and changes in the feed to milk price ratio. The information gathered will be used to develop recommendations for successful use of molasses on organic dairies to be disseminated to the industry. Key words: dairy, molasses, supplementation 203 14 POTENTIAL OF LEGUMES AS SUBSTITUTES FOR NITROGEN FERTILIZER IN SUMMER STOCKER GRAZING SYSTEMS. R. R. Reuter*1, J. T. Biermacher1, J. K. Rogers1, T. J. Butler1, M. K. Kering1, J. R. Blanton1, and J. A. Guretzky2, 1The Samuel Roberts Noble Foundation, Inc., Ardmore, OK, USA, 2The University of Nebraska-Lincoln, Lincoln, NE, USA. Considerable interest exists among producers concerning the biological and economic impacts of replacing synthetic nitrogen fertilizer with legumes in bermudagrass pastures in the southern Great Plains. The objective of this experiment was to compare stocker cattle performance and net returns of two alternative legume systems in bermudagrass with the conventional practice of applying fertilizer. Therefore, 9, 1.4-ha bermudagrass (Cynodon dactylon) paddocks were randomly assigned to one of three management systems in a completely random design in 2007. The first management system was nitrogen fertilizer (BG/N; 112 kg/ha actual nitrogen). The second system was bermudagrass interseeded with a perennial legume [BG/A; alfalfa (Medicago sativa)]. The third system was bermudagrass interseeded annually with an annual legume mix [BG/C; hairy vetch (Vicia villosa), crimson clover (Trifolium incarnatum), and arrowleaf clover (Trifolium vesiculosum)]. Legumes were drilled into the bermudagrass sod in the fall. Alfalfa did not establish successfully the first year, and was replanted in the fall of 2008. These paddocks were grazed for two years (2008 and 2009) with 238-kg steers stocked at 2.9 steers / ha during the spring and summer as long as forage availability exceeded 1121 kg/ha. Steers were shrunk in drylot and weighed every 28 d during grazing. Data were analyzed as a mixed model CRD with 3 replications in SAS. The model included a fixed effect of management system and a random effect of year for all dependent variables. Least squares means of management systems were separated by least significant difference in the presence of a significant F-test. Average grazing days were greater for BG/N than for BG/C and BG/A (110, 67, and 76 d, respectively, P < 0.05). Forage availability (measured by hand clipping every 14 d during grazing) was also greater for BG/N as compared to BG/C and BG/A (2374, 1745, and 1520 kg/ha respectively, P < 0.05). Forage nutritive value was determined by NIRS analysis of 14-d, hand-clipped samples. Forage IVDMD percentage was greatest for BG/C and least for BG/N, with BG/A intermediate (73.7, 69.4, and 71.9, P < 0.05). Further, BG/C CP percentage was greater than either BG/N or BG/A (16.9, 14.6, and 14.4, P < 0.05). Interestingly, these improvements in forage nutritive value did not translate into greater steer ADG (0.80, 0.82, and 0.76 kg/d for BG/N, BG/C, and BG/A systems, respectively; P > 0.05). However, body weight gain / ha favored (P < 0.05) the BG/N system by an average of 74 kg/ha over the legume systems. Total production cost for the BG/N, BG/C and BG/A systems averaged $188, $215, and $114/ha, respectively (seed cost in the BG/A system was prorated across an expected 5 year life of the alfalfa stand). Recurring annual seed costs coupled with relatively low steer gain / ha caused net return for the BG/C system to be negative in both years. During this study, actual nitrogen fertilizer cost for the BG/N system was $0.99 / kg of actual nitrogen. The nitrogen cost that would cause the BG/A system to breakeven with the BG/N system was $1.35 / kg. In 2008, the price of nitrogen fertilizer approached $2.00 / kg in some regions of the southern Plains. Therefore, in situations of very high nitrogen fertilizer prices, a perennial legume may be economically competitive with synthetic fertilizer for use in grazed bermudagrass pastures in the southern Great Plains. Total production cost and gain / ha (largely a function of grazing days) were the primary variables that influenced net return. Legume effects on animal performance were of relatively minor importance. Key words: bermudagrass, legume, fertilizer 204 15 EFFECTS OF CONVENTIONAL AND GRASS FED SYSTEM ON CARCASS TRAITS AND PERFORMANCE OF ANGUS STEERS. G Cruz*, G Acetoze, and H Rossow, Univerisity of California, Davis, CA USA. Although the US market for grass finish beef is small, the demand has increased. Beef producers have been working on different management, nutrition and breeding strategies to increase the sustainability of this system. The objectives of our study were to quantify differences in daily gain and carcass characteristics for Angus steers fed a high concentrate diet or only forage to identify steers that would perform well in a grass finish system. Thirty animals were randomly selected from a local ranch. Fourteen were fed a corn based finishing diet (10.0% CP, 20% NDF) at the University of California, Davis feedlot and the remaining stayed on a white clover and rye grass (15.8 %CP and 49.8% NDF) pasture. Animals were weighed and ultrasound once a month. For both feedlot and grass finish groups, cattle were fed until they reached similar quality grade (Se+, Ch-) at time of slaughter using ultrasound. Differences in measures of performance, ultrasound and carcass merit among steers in different dietary treatments were determined using the linear procedure in R Programming, and included the fixed effect of dietary treatments. Mean initial body weight of the two groups were not statistically different with 414.5 kg and 409.7kg (P=0.8) for feedlot and grass-finish, respectively. As expected, animals at the feedlot had higher average daily gain (1.35 vs. 0.73 kg/d; P<0.001), rib-eye area gain per day (0.08 vs. 0.03 cm2/d; P<0.05), back fat gain per day (0.004 vs. 0.001 cm/d; P<0.001) and rump fat gain per day (0.005 vs. 0.002 cm/d; P<0.001).The average days on feed for the feedlot animals were significantly lower than pasture animals (167 vs. 286 days; P<0.001). When comparing carcass characteristics, feedlot animals had a higher hot carcass weight (404 vs. 352 kg; P<0.001), dressing percentage (63 vs. 57%; P<0.001), Yield Grade (3.5 vs. 2.9; P<0.001) and Quality Grade (Ch vs. Se+; P<0.001). Feedlot steers also had a higher back fat (1.19 vs. 0.89 cm; P<0.001) and rump fat (1.39 vs. 1.06 cm; P<0.001), yet rib eye areas were similar (75.5 vs. 76.3 cm2; P=0.8) which explains the better Quality Grade results. Overall, feedlot steers gained more, were larger and had more fat cover over a shorter period of time than the forage finished steers. However, grass finish steers were able to have large rib eye areas and finish on forage if given enough time and quality forage. Therefore criteria for evaluating carcasses from grass and feedlot finish steers should be different, and this process will help develop breeding strategies to select animals that will better perform in grass finish systems. Key words: beef cattle, grass fed, carcass traits 205 16 ASPECTS OF CARCASS COMPOSITION AND ORGAN WEIGHTS OF ANGUS STEERS FINISHED ON GRASS AND HIGH GRAIN DIETS. G. Acetoze*, G. D. Cruz, and H. A. Rossow, University of California, Davis, California, USA. Interest in grass finished beef has been increasing. However more research is needed to apply grass finishing to current beef production systems. A common method to predict carcass composition is based on the composition of 9-11th rib section. However, this method was developed with data from grain finished cattle. In addition, metabolic rate and therefore energy requirements are closely linked to organ sizes of metabolically active organs such as liver, gut and kidney. The objective of this study is to examine the effects of strictly forage or high grain diets on organ weights and carcass composition of Angus steers. Carcass composition was estimated using the composition of 911th rib section from the right side, to compare the percentages of muscle, fat and bone of 14 grain-fed and 13 pasture-fed Angus steers. All internal organs were weighed and sampled to compare estimates of basal metabolic rate among feedlot and grass finish cattle. Steers were slaughtered when their estimated quality grade by ultrasound was greater than low select. Average yield and quality grades for grass finish steers were 58% and high select, and 63% and average choice for grain finish steers. Pasture was a mix of white clover and ryegrass (15.8% crude protein, 49.8% NDF) and the feedlot diet was a 90% corn, 10% alfalfa finishing diet (10.0 % crude protein, 20% NDF) on a dry matter basis. Data were analyzed with the general linear models procedure of SAS (SAS Institute, 2004). Muscle and fat % were significantly different (P< 0.01) among diet groups with means for muscle of 56.84% and 61.46% and fat of 24.66% and 20.70% for grain finish and grass finish, respectively. Bone % was not significantly different (P > 0.01) and the means were 14.17% and 15.82% for grain finish and grass finish, respectively. The following organs were not statistically different (P > 0.01) and their averages were: spleen 1.14 kg (SE = 0.062) and 1.06 kg (0.041), liver 7.04 kg (0.196) and 6.83 kg (0.166), kidney 0.52 kg (0.016) and 0.55 kg (0.017), gallbladder 0.41 kg (0.041) and 0.49 kg (0.034), heart 2.66 kg (0.074) and 2.56 kg (0.083), lungs 8.65 (0.425) and 8.83 kg (0.232), for grain finish and grass finish, respectively. These results show that grass and grain finish diets differ in carcass composition but organ weights are not significantly different. Key words: grass finish, carcass composition, organ weight 206 17 PERFORMANCE AND EFFICIENCY ON PASTURE OF CATTLE WITH DIVERGENT PHENOTYPES FOR RESIDUAL FEED INTAKE. T.D.A. Forbes*1, A.D. Aguiar2, L.O. Tedeschi3, F.M. Rouquette, Jr.4, G.E. Carstens3, and R.D. Randel4, 1Texas AgriLife Research, Uvalde, TX, USA, 2University of Florida, Gainsville, FL, USA, 3Texas A&M University, College Station, TX, USA, 4Texas AgriLife Research, Overton, TX, USA. The objectives were to 1) examine performance and efficiency on pasture of cattle with divergent phenotypes for residual feed intake (RFI) and to compare RFI obtained under grazing (RFIg) with previously determined RFI obtained under confinement feeding (RFIc) using Braunveih-sired crossbred steers and Brahman bulls; 2) assess the feasibility of determining DMI of Brahman bulls on Coastal bermudagrass [Cynodon dactylon (L.) Pers.] pasture using alkane-marked corn gluten individually-fed via Calan gates; and 3) compare the use of bermudagrass forage C31 or C33 alkane in the DMI calculation. Braunveih steers (n = 169) were fed a high roughage diet (ME = 2.2 Mcal/kg DM) in Calan gates for 77 d, and RFIc calculated as the residual from the linear regression of DMI on midtest BW0.75 and ADG. Steers were ranked by RFIc, and those with low (LRFI), medium (MRFI) and high (HRFI; n = 9, respectively) RFI were subsequently placed on annual ryegrass (Lolium multiflorum) pasture for an 84-d grazing trial. Forage DMI was determined on two separate occasion s using continuous release capsules containing C32 and C36. Fecal samples were collected twice daily for 10 d, and forage samples collected daily for analysis of n-alkane concentrations via gas chromatography. Brahman bulls (n=53) were fed a high roughage diet for 70 d in Calan gates, and bulls with low and high RFIc (n = 8) allotted to 4 replicate Coastal bermudagrass pastures (2 low and 2 high RFI bulls per pasture) and grazed for 56 d. Bulls were individually fed twice daily with 400 g corn gluten marked with C32 for four 10-d periods, using Calan gates in pasture. Fecal samples were collected 4 times daily (0700, 1100, 1500, and 1900 h) for the last 5 d of each period, and forage samples collected over the last 7 d to determine forage n-alkane concentrations. Braunveih steers with LRFI gained similarly (P = 0.26) to MRFI and HRFI steers (1.10, 1.06, and 1.00 ± 0.04 kg/d, respectively), and DMI did not differ (P = 0.28; 6.4, 6.4, and 7.0 ± 0.28 kg DM/d). Despite lack of difference in forage DMI, RFIc was moderately correlated (P < 0.05) with RFIg (r = 0.39), and the ratio of forage DMI to ADG was lower (P = 0.001) in LFRI and MRFI than HRFI (5.8, 6.0, and 7.0 ± 0.28 respectively). In the Brahman bull study, four calculations (CALC) were used to estimate DMI: C31 or C33 with adjustment for forage C32 (C31 and C33), and C31 and C33 without adjustment for C32 (C31_0 and C33_0). Within a CALC, treatments (TRT) were assumed to be either individual or combination of different times of daily fecal collection. Statistical analyses included the fixed effects of TRT and RFIc, and the random effects of period, d within period, animal, and pasture. Bulls did not consistently consume all the corn gluten in period 1 which was dropped from further analysis. Gain did not differ between HRFI and LRFI bulls while grazing. Predicted DMI (kg/d) using C31 (9.7), C33 (6.4), C31_0 (8.9), and C33_0 (5.9) alone or in combination [C31 and C33 (7.4), or C31_0 and C33_0 (7.4)] differed (P = 0.011). Morning and afternoon fecal collections differed (P < 0.0010) in predicting DMI with C31 (9.44), C33 (6.29), C31_0 (8.68), and C33_0 (5.74) for the morning and C31 (10), C33 (6.66), C31_0 (9.20), and C33_0 (6.09) for the afternoon. There were no differences in predicted DMI between Brahman bulls ranked as efficient or inefficient using any n-alkane CALC (P > 0.90). A nonparametric analysis indicated that pre-ranking animals for RFI under confinement did not predict (P < 0.0001) similar RFI rankings in the Brahman bulls. These results suggest that using the combination of Calan gates, alkane-marked corn gluten and twice daily fecal collection (0700 and 1500 h) it is possible to accurately estimate DMI of cattle grazing Coastal bermudagrass pastures. Further studies are planned to explore possible explanations for the correlation between RFIc and RFIg in the Braunveih steer study and the lack thereof in the Brahman bull study. Key words: cattle, efficiency, residual feed intake 207 18 EFFECTS OF SAND SAGEBRUSH CONTROL ON STOCKER CATTLE PERFORMANCE ON A SOUTHERN MIXED PRAIRIE COMMUNITY. S. A. Gunter*1, T. L. Springer1, E. Thacker1, and R. L. Gillen2, 1 USDA-ARS, Southern Plains Range Research Station, Woodward, OK USA, 2Western Kansas Agricultural Research Centers, Kansas State University, Hays, KS USA. Research on sand sagebrush communities has focused primarily on sagebrush control methods, with minimal attention to effects on livestock production. To evaluate the effects of sand sagebrush (Artemisia filifolia) control on cattle production, 15 native pastures (10 to 21 ha each) were selected in 2003 at the Southern Plains Experimental Range near Ft. Supply, Oklahoma. Five of the pastures had recently (< 5 yr) been sprayed with 2,4 dichlorophenoxyacetic acid to control sand sagebrush, 6 had received chemical treatment before 1994, and the remaining 4 had received no treatment. The 15 pastures were categorized into 3 discrete sagebrush levels based on past management: 1) High with no sagebrush control within the last 60 yr, 2) Medium with sagebrush control at least 9 yr and older, and 3) Low with sagebrush sprayed in 2003 with 1.1 kg of 2,4 dichlorophenoxyacetic acid (a.i.)/ha. The average sagebrush canopy cover from 2003 to 2008 in the High, Medium, and Low pastures was 26.4 (n = 4), 10.0 (n = 6), and 3.0% (n = 5), respectively, as determined annually by the line-transect method. The major grass species in the pastures were blue grama (Bouteloua gracilis), sand dropseed (Sporobolus cryptandrus), little bluestem (Schizachyrium scoparium), sand lovegrass (Eragrostis trichodes) and sand bluestem (Andropogon hallii). From 2004 to 2008, approximately 106 steers (BW = 202 ± 6.1 kg) were stocked annually in late January and subsequently removed in mid-August. High pastures were stocked at 47 animal-unit-d/ha and Medium and Low pastures were stocked at 69 animal-unit-d/ha. Cattle were supplemented with oil seed meal based cubes (48% CP) from January through mid-April at a rate of 0.68 kg/steer daily; no supplement was fed from April to August. Mineral blocks (NaCl) were available during the entire trial. Initial BW in January did not differ (P = 0.57) among treatment (203, 201, and 202 kg for High, Medium, and Low, respectively). Also, BW did not differ (P = 0.76) in April (236, 235, and 235 kg for High, Medium, and Low, respectively, or at removal in August (P = 0.12; 352, 344, and 346 kg for High, Medium, and Low, respectively). Average daily gains from January to April did not differ (P = 0.96), but did differ from April to August (P = 0.02). Steers grazing the High pastures gained BW faster (P < 0.01; 1.03 kg/d) than steers grazing Medium pastures (0.96 kg/d) and tended to gain faster (P = 0.08) than steers grazing Low pastures (0.98 kg/d). Body weight gain per steer during the January to April grazing period did not differ (P = 0.80). However, steers in the High pasture had higher BW gain per steer from April to August (P < 0.01) than steers in the Medium pastures. Steers in the Medium pastures tended (P = 0.08) to have higher BW gain per steer than the Low pastures. Even with the differences in BW gain per steer during the summer, the overall BW gain per steer did not differ (P = 0.14). Body weight gain per hectare was greatest (P < 0.01) for Low pastures (102 kg) compared to High (67 kg) or Medium (81 kg) pastures; also, Medium pastures were greater (P < 0.01) than High pastures. No sagebrush control (High), with an appropriate stocking rate resulted in the lowest BW gain hectare. However, the use of intensive sagebrush control (Low; spraying every 5 yr) and increasing the stocking rate (47%) over the High pastures resulted in a 26% increase in BW gain per hectare. However, the moderate sagebrush control (Medium; spraying intervals of approximately 10 yr) resulted in performance similar to that of the Low pastures. These results suggest that medium levels of sand sagebrush control with appropriate stocking rates may be the best management option for stocker cattle production. Key words: cattle, herbicides, sand sagebrush 208 19 PASTURE MANAGEMENT EFFECTS ON NONPOINT SOURCE POLLUTION OF MIDWESTERN PASTURES. D.A. Bear*, J.R. Russell, and D.G. Morrical, Iowa State University, Ames, Iowa, United States of America. Grazing management practices that allow cattle to congregate near pasture streams may increase sediment, nutrient, and pathogen loading of the streams by removing vegetation and depositing manure in and near the streams. To assess effects of microclimate conditions and the size and/or shape, botanical composition, and shade distribution of pastures on the congregation of cattle near pasture streams, grazing distribution of cattle on five beef cow-calf farms in southern Iowa were analyzed during spring, summer, and fall of 2007 through 2009. Pastures ranged in size from 8 to 125 ha with varying proportions of cool-season grasses, legumes, weeds, brush, and bare ground. Cows were Angus and Angus-Cross on four of the farms, and Mexican Corriente on the remaining farm. Two to 3 cows on each farm were fitted with Global Positioning Systems (GPS) collars to record location within a pasture at 10 min intervals for periods of 5 to 14 d in each season. HOBO data loggers recorded ambient temperature, black globe temperature, dew point, relative humidity, and wind speed and direction at 10 min intervals during the three-year project. Data were used to calculate 13 heat indices to be used to estimate cow distribution in or within 30.5 m (streamside zone) of a water source within pastures. Water sources and fence lines were referenced on a geospatial map and used to establish zones within the pastures. Designated zones were in the stream or pond (water source) or lower or greater than 30.5 m (Uplands) from the water source. One hundred thirty-nine data sets were obtained during the three grazing seasons. Farm and seasonal effects on cow distribution in pastures were analyzed with the GLM procedure of SAS using years as replicates. Mean proportions of observations when cattle were in the water source differed (P < 0.01) between seasons being 1.1, 1.8, and 1.0% in the spring, summer, and fall, respectively. Mean proportions of observations when cattle were within the streamside zone did not differ (P > 0.05) between seasons, but differed (P < 0.01) between farms. Heat indices effects on the probability of cattle being in or within 30.5 m of the water were estimated using the LOGISTIC procedure of SAS. Ambient temperature best predicted cattle location in the streamside zone compared to the other twelve indices, and determined the probability of finding cattle in the streamside zones of pastures by farms ranged from 2.6 to 4.9% for every 1 Celsius degree increase in ambient temperature. The proportion of time cattle were within the streamside zone increased with decreasing pasture size (y = 32.88 - 0.79x + 0.0054x2, r2 = 0.49; y = 35.73 - 0.73x + 0.0043x2, r2 = 0.77; y = 36.49 0.89x + 0.0058x2, r2 = 0.66), increasing the proportion of streamside zone within a pasture (y = -3.05 + 2.34x 0.059x2, r2 = 0.47; y = 5.75 - 0.32x + 0.075x2, r2 = 0.37; y = 0.70 + 0.93x - 0.0094x2, r2 = 0.39), and increasing the proportion of shade located within the streamside zones of pastures (y = 1.76 + 0.89x - 0.011x2, r2 = 0.36; y = 7.08 + 0.40x - 0.0031x2, r2 = 0.25; y = 2.11 + 0.82x - 0.0091x2, r2 = 0.31) in the spring, summer, and fall, respectively. Botanical composition of pastures were evaluated and regressed on cattle locations within the streamside zone, but no relationship existed. Consequently, the proportion of time cattle were located within the streamside zone increased with increasing ambient temperature, decreasing pasture size, increasing the proportion of streamside zone within a pasture, and proportion of shade located within the streamside zone. Implementation of grazing management practices, such as implementation of a rotational grazing system, flash grazing of the streamside zones of pastures, altering pasture size, shape, and shade distribution within a pasture altering the location, timing, intensity, and length of grazing for the protection of pasture streams will likely be most effective on small and/or narrow pastures in which cattle have less opportunity to locate in upland locations. Key words: environment, grazing behavior, water quality 209 20 SUPPLEMENTATION OF RUMINALLY UNDEGRADABLE PROTEIN TO MAINTAIN ESSENTIAL AMINO ACID SUPPLY DURING NUTRIENT RESTRICTION ALTERS CIRCULATING ESSENTIAL AMINO ACIDS OF BEEF COWS IN EARLY TO MID-GESTATION. A. M. Meyer*1, J. S. Caton1, M. Du2, and B. W. Hess2, 1Center for Nutrition and Pregnancy, Department of Animal Sciences, North Dakota State University, Fargo, ND, USA, 2Department of Animal Science, University of Wyoming, Laramie, WY, USA. Early to mid-gestation is a critical period for placental growth and fetal organogenesis, thus maternal nutritional insults during this period may be particularly detrimental to offspring development. We hypothesize that supplementation of ruminally undegradable protein (RUP) to provide essential AA will offset negative consequences of maternal nutrient restriction on offspring. Our specific objective was to determine the effect of RUP supplementation on serum essential AA concentration in beef cows nutrient restricted during early to midgestation. Angus Gelbvieh cows (n = 36) were blocked by parity and randomly allocated by BW to 1 of 3 dietary treatments for 140 d, beginning on d 45 of gestation. Treatments included a native grass hay-based control (CON) diet formulated to meet NRC requirements for mid-gestation; a nutrient restricted (NR) diet providing 70% of CON NEm ; or an NR diet fed with a RUP supplement (NRP; 68.7% menhaden fish meal, 24.5% hydrolyzed feather meal, and 6.8% porcine blood meal, DM basis) formulated to provide similar essential AA flow to the duodenum as CON. Cows were individually fed, and diets were adjusted biweekly for BW change and increasing NEm requirements of gestation. Blood samples were obtained and serum collected on d 0, 42, 84, and 140 of the treatment period for AA analyses. Data were analyzed using a mixed model containing effects of treatment, sampling day, and their interaction, where sampling day was used as a repeated effect. There were treatment main effects for total serum AA (P = 0.002), His (P = 0.001), Lys (P = 0.06), and Trp (P = 0.10). Total AA were greater (P < 0.006) for cows fed NRP compared with CON and NR (2.48 vs. 2.3 and 2.25 ± 0.04 mmol/L). Serum His was less (P < 0.002) for CON than NR and NRP, whereas CON had greater (P < 0.04) Lys and Trp than NR. A treatment sampling day interaction was detected (P 0.08) for total essential AA, all other essential AA (mol/L), and all relative essential AA (% of total essential AA) concentrations. On d 42 of the treatment period, cows fed NRP had greater (P < 0.001) total essential AA, Arg, Val, Leu, and Phe than CON and NR, whereas NR had less (P < 0.05) Ile than CON and NRP. Cows fed NRP continued to have greater (P < 0.04) serum Leu and Phe on d 84. Total essential AA, Arg, and Val on d 84 and 140 and Leu on d 140 were greatest (P < 0.03) for cows fed NRP, intermediate for CON, and least for NR. On d 84 and 140, NR cows had less Ile (P < 0.02) compared with CON and NRP. Cows fed the CON diet had greater (P < 0.007) Met than NR and NRP on d 140, whereas NRP had greater (P = 0.005) Phe than NR. When expressed as a percent of total essential AA, NRP cows had less (P 0.003) Thr, Met, and Trp from d 42 to 140, Lys on d 42 and 84, Arg on d 42, and Ile on d 140 than CON and NR. Cows fed the NR diet had greater (P < 0.001) His than CON and NRP from d 42 to 140, whereas NR had greater (P < 0.009) Phe than NRP on d 42 and 140. Serum Ile was greatest (P < 0.005) for CON, intermediate for NR, and least for NRP on d 42 and 84. Cows fed the CON diet also had greater (P = 0.001) Lys than NRP on d 140. Conversely, NRP had greater (P < 0.001) Val and Leu compared with CON and NR from d 42 to 140. Providing nutrient restricted beef cows with RUP during gestation can increase serum essential AA, but may also alter their relative concentrations. Increasing circulating essential AA in nutrient restricted cows may protect the developing fetus from intrauterine growth restriction, although more research is necessary to determine these effects. Key words: amino acids, gestation, nutrient restriction 210 21 EFFECT OF FORAGE TYPE ON PERFORMANCE OF WEANLING BEEF STEERS. G. Scaglia*1, B. Corl2, W. S. Swecker, Jr.3, and A. Lillie4, 1Louisiana State University Agricultural Center, Iberia Research Station, Jeanerette, LA, USA, 2Department of Dairy Science, Virginia Polytechnic Institute and State University, Blacksburg, VA, USA, 3Virginia Maryland Regional College of Veterinary Medicine,Virginia Polytechnic Institute and State University, Blacksburg, VA, USA, 4Department of Animal and Poultry Sciences, Virginia Polytechnic Institute and State University, Blacksburg, VA, USA. In two consecutive years, 42 weanling crossbred Angus steers (BW = 266 ± 5.2 kg) were purchased at auction and transported to the Virginia Tech Kentland Farm, VA. The experimental design was a completely randomized design with 3 replicates. In each block there were 6 paddocks and of those, three paddocks were planted with each of the two forages used. Three paddocks were planted (September 2, 2004) with Jesup endophyte-free tall fescue (Lolium arundinaceum (Schreb.) S.J. Darbyshire = Festuca arundinacea Schreb) at a rate of 11.3 kg/ha and the other three paddocks were planted (September 5, 2004) with AmeriStand 403 alfalfa (Medicago sativa) variety at a rate of 9 kg/ha. Tall fescue pastures received 68 kg/ha of diammonium phosphate (DAP) and the alfalfa pastures received 68 kg/ha of DAP + 45.5 kg/ha of Potash . Nitrogen was applied to the tall fescue paddocks at 13.6 kg/ha in October (2004) and then again in March (2005) to promote vigorous growth and tillering. Each year 42 weanling steers were blocked by weight into 6 groups of 7 steers each and randomly allotted and rotationally stocked to alfalfa or tall fescue within block, where they would remain until the end of the grazing period. All paddocks had an area of 0.61 ha (stocking rate of 3.8 steers/ha) . The steers grazing the tall fescue paddocks were provided with trace mineralized salt. The mineral blocks in the alfalfa pastures contained poloxalene in addition to the trace mineral in order to help prevent bloat. All steers were weighed on d 0, 14, 28 and every 28 d thereafter until the conclusion of the grazing period. Cattle rotation between paddocks depended upon forage availability. When the tall fescue was grazed to a height of approximately 5 cm the cattle were moved to the next paddock. When the alfalfa was grazed to a height of 10 to 15 cm cattle were moved to the next paddock. Forage mass for the fescue and alfalfa paddocks were determined by measuring and cutting 18 double samples using wire circles of 0.25 m2 at 2.5 cm above ground level. Additionally 30 measurements using a falling plate meter to determine canopy height per paddock were taken. Nutritive value samples were taken on d 0 and every 14 d thereafter. Nutritive value samples were collected by walking two diagonal strips in an X pattern in the designated paddock and grab samples were obtained approximately every ten steps. Data was analyzed using PROC MIXED of SAS. For data analysis period was used as repeated measures and the experimental unit was pasture within treatment (replicate). Treatment was the fixed effect and year the random effect. The model included treatment, replicate, period, and treatment x period interaction. In all cases level of significance was declared at P < 0.05. Trends were declared at 0.05 P0.10. Forage mass was different among years (P = 0.002), sampling day (P = 0.023) and treatments (P = 0.006). No interactions were detected (P > 0.05) for forage mass determination. Less forage was produced in year 1 than in year 2 (2526 vs. 3244 kg DM/ha for weanlings, in year 1 and 2, respectively). Both forage stands were new stands so it is normal to see an increase in forage mass in the second year of production. Alfalfa had greater DM production (3203 kg DM/ha) than tall fescue (2568 kg DM/ha). A possible explanation for these results is that the alfalfa variety used in this experiment was selected for grazing; hence we can expect higher DM production and resistance to grazing. Forage mass was lowest (2359 kg DM/ha) on d 56 and different (P < 0.05) than on d 28 (3387 kg DM/ha) and 84 (3133 kg DM/ha). As expected, alfalfa presented a larger concentration of CP but lower NDF and ADF than tall fescue at all sampling times (P < 0.01). The interaction of treatment x day of sampling was significant (P = 0.001) for NDF and ADF. There is an increase on NDF and ADF concentrations from d 0 to d 84 for tall fescue and to d 112 for alfalfa, declining thereafter. The differences in ADG between d 28 and 84 explained the ADG differences for the whole grazing period (d 0 to 140) as well as the differences in final BW and in kg of beef produced per unit of area (kg/ha). Steers grazing alfalfa gained 0.37 kg/d more than those grazing tall fescue (0.99 vs. 0.62 kg/d). It resulted in 332 and 524 kg/ha of beef produced for tall fescue and alfalfa, respectively. Even though alfalfa is known to be a better forage, steers grazing tall fescue gained adequately. Key words: alfalfa, tall fescue, steers 211 22 ESTIMATES OF DRY MATTER INTAKE OF BEEF STEERS GRAZING HIGH QUALITY PASTURES USING ALKANES. G. Scaglia*1 and H. Boland2, 1Louisiana State University Agricultural Center, Iberia Research Station, Jeanerette, LA, USA, 2Prairie Research Unit, Mississippi State University, Prairie, MS, USA. Pastures remain the most important source of nutrients for ruminant livestock and nutrition is critical to optimize animal production. The daily quantity of dry matter that is consumed by an animal is a critical measurement to make nutritional inferences about feed and subsequent animal response. Researchers are facing the dilemma that, while estimates of individual animal performance are readily obtained, it is still difficult to estimate the herbage intake of individual animals. The objectives of this experiment were to estimate forage intake in beef steers grazing tall fescue [Lolium arundinaceum (Schreb.) Darbysh. = Festuca arundinacea Schreb.] or alfalfa (Medicago sativa)/tall fescue mixed pastures and to measure the recovery rate of artificial alkanes from a controlled release device under these conditions. Six steers (330±11 kg) were allotted to each pasture (0.61 ha). Steers were previously trained to the use of a harness and fecal collection bags, so that the expected negative effect on animal behavior could be minimized. A controlled-released capsule (Nufarm, Auckland, NZ) containing alkanes (C32 and C36; release rate: 400 mg/d) was administered on d 0 to all 12 steers. During the experimental period (d 8 to 14) fecal collection bags were fitted to all steers in the fescue pasture and to three steers in the fescue/alfalfa pastures. Collection bags were emptied twice daily (at 0830 and 1630). Feces were weighed, mixed, and an aliquot of 0.5 kg frozen. Rectal grab samples were obtained twice daily from each animal at the time of switching bags. Forage mass in each pasture was estimated from two 3 meters strips obtained on d 8 using a push mower. From d 6 to 12 forage samples for alkane determination were obtained by walking the pasture in an X and clipping every 20 steps with a set of handheld clippers at a height of 2.54 cm from the ground. Fecal and forage samples were freeze dried and ground through a 0.5 mm screen. Alkane determination and dry matter intake estimation used C31 as odd chain alkane following the method. The effect of pasture on herbage intake was estimated by analyses of variance. A paired t-test was used to compare herbage intake estimated from grab samples vs. samples taken from the fecal collection bag. All statistical analyses were conducted using SAS. Forage availability (3078 vs. 3088 kg DM) and quality (CP: 19 vs. 17.5%; NDF: 50 vs. 55%; ADF: 31 vs. 29%) were similar for fescue/alfalfa and fescue pastures, respectively. Similar concentration of C32(mg/kg of feces) was found in fecal samples taken directly from the steers (grab samples) and from the fecal collection bags, with an average ratio (C32 grabs/ C32 bags) of 0.99 and 0.92 for alfalfa/fescue and fescue pastures, respectively. The recovery rate of C32 in feces was higher than previously reported with an average of 0.985 and 0.991 for fescue/alfalfa and fescue pastures, respectively. Using information obtained from fecal samples collected from bags there was no difference (P > 0.05; SEM=0.48) in the average daily dry matter intake of steers grazing fescue/alfalfa (10.8 kg DM) and fescue pastures (10.3 kg DM). Similarly, no differences (P >0.05) were detected in average daily dry matter intake when using C32 from grab samples (9.9 vs. 9.6 kg DM for steers grazing fescue/alfalfa and fescue pastures, respectively; SEM=0.39). When sampling method was compared (grab fecal samples vs. fecal samples from bags) there was no difference (P >0.05) in the estimation of dry matter intake for steers grazing fescue or fescue/alfalfa pastures. The results of the present study demonstrate that knowing the recovery rate of the marker used (in this case C32) fecal grab samples can be used to estimate dry matter intake in beef steers under grazing conditions. However, the recovery rate for C32 observed in this experiment was 5 to 15% higher than those found in the literature. Therefore, further research is needed to study the effectiveness of the controlled-release capsules under different conditions and factors that might be affecting the release rate of the marker. Key words: alkanes, intake, pastures 212 23 EFFECT OF ESTABLISHMENT METHOD OF WHEAT PASTURE AND FALL STOCKING RATE ON PERFORMANCE OF GROWING STEERS. P Beck*, M Morgan, T Hess, D Hubbell, M Anders, and B Watkins, University of Arkansas Division of Agriculture, Little Rock, AR USA. Agricultural practices have impacted the environment for many centuries. In past generations, agricultural disasters like the Dust Bowl in the 1930’s increased awareness about tillage practices and their impact on the environment. Passage of the Clean Water Act in 1972 set in motion actions that have greatly impacted land use by farms. Previous research concerning the impact of tillage systems on environmental health has focused on crop production with little research to determine the interaction of livestock grazing and tillage system on resource sustainability of farms. Rainfall is critical for forage production and can be an important concern when either inadequate or overabundant. Dry conditions limit forage establishment and growth while wet conditions can delay establishment due to crusting of the soil surface, will cause footing problems for livestock, and hoof action of livestock will disturb wheat stands. This research was conducted over 3-yr from fall of 2006 to spring of 2009 to determine the effect of fall stocking rate (SR) of small grain pastures on animal performance during the fall and subsequent spring for 3 establishment methods: 1) no-till (NT) seeding into undisturbed stubble with approximately 85% residue cover, 2) reduced-till (RT) - disking once followed by broadcast seeding with approximately 50% residue cover, and 3) conventional-till (CT) – drilling into a prepared seedbed with approximately 4% residue cover. Each year soft-red winter wheat was sown in the first wk of September (136 kg/ha). During the fall 60 steers (BW = 226 ± 17.9 kg) were stocked to pastures at 1.8, 2.4, and 3.6 steers/ha (n = 2 pastures/stocking rate and establishment method) and grazed from early November until mid-February (depending on forage availability). Each spring graze-out period 180 steers (BW = 246 ± 21.0 kg) were stocked to pastures at 10 steers/ha in late February or early March (depending on forage availability) and removed in early May. Animal performance was analyzed with regression using SAS; indicator variables were used to construct orthogonal contrast in order to separate the effects of tillage (NT vs RT and CT) and CT vs RT. There was no establishment method effects (P 0.24) on fall BW gain, but fall BW gain decreased linearly with increasing SR (Total BW gain, kg = 166 18*SR, R2 = 0.24, P < 0.001). Effect of fall SR on spring graze-out BW gain and ADG were quadratic (P < 0.01), with interactions (P < 0.05) present for tillage system comparisons with NT BW gain = 162 - 91.5*SR + 16.2*SR2 (P < 0.01, R2 = 0.11), CT BW gain = 160 - 85.9*SR + 13.9*SR2 (P < 0.01, R2 = 0.13), and RT BW gain = 125.8 – 52.8*SR + 7.3*SR2 (P < 0.01, R2 = 0.21); and NT ADG = 3.50 – 2.01*SR + 0.37*SR2 (P < 0.01, R2 = 0.13), CT ADG = 3.40 – 1.88*SR + 0.32*SR2 (P < 0.01, R2 = 0.10), and RT ADG = 2.46 – 0.98*SR + 0.14*SR2 (P = 0.05, R2 = 0.15). Total BW gain/ha of NT increased in a quadratic (P < 0.01, R2 = 0.80) fashion with increasing SR (BW gain/ha = 1040 – 438*SR + 96*SR2). Stocking rate during the fall did not affect (P > 0.50) gain/ha of RT (612 kg/ha) or CT (579 kg/ha), neither of which differed (P > 0.70) from NT. Reduced tillage tended (P = 0.07) to have greater BW gain/ha than CT. Results from this study indicate that tillage system has no affect on animal performance during the fall. Although animal performance is reduced in the fall grazing period with increasing SR, gains would still be considered adequate for most producers at the greatest SR. Increasing SR in the fall reduces animal performance the following spring, except for NT, indicating there was less impact on forage production during the spring with NT than with CT or RT. Contrary to conventional wisdom there was no difference in BW gain/ha between CT and NT or RT. Selection of an establishment method is ultimately dependent on economic conditions and producer’s sensitivity to environmental impacts. Key words: tillage system, growing cattle, stocking rate 213 24 GROWTH PERFORMANCE EFFECTS OF MINERAL VS. SALT SUPPLEMENTATION ON STOCKER CATTLE GRAZING SPRING-SUMMER PASTURES OVER TWO CONSECUTIVE YEARS. P. J. Guiroy*, S. E. Showers, and B. McMurry, Cargill Inc., Hopkins, MN. The objective was to evaluate growth performance in yearling crossbred steers grazing native spring-summer pastures in Osage County, Oklahoma for the effects of mineral vs. salt (NaCl) supplementation over two consecutive years. Year 1 consisted of 2,132 steers grazed for 85 days (mid April to mid July 2008). Year 2 consisted of 2,225 steers grazed for 123 days (late March to mid July 2009). Each year animals were randomly assigned to one of 8 pasture paddocks, 4 paddocks (replicates) per treatment. The same pastures were utilized in years 1 and 2, with the mineral treatment paddocks in year 1 becoming the salt treatment paddocks in year 2. The mineral used in this study (NutreBeef® Stocker Summer Mineral) contained a minimum of 18% calcium, 8% sodium, 4% phosphorus, 1% magnesium, trace minerals and vitamins, and was formulated to provide 350 mg/hd/day of Chlortetracycline. Data obtained from this experiment were analyzed by analysis of variance using Proc GLM as complete randomized design. The model included year, treatment, and the interaction year x treatment effects, with pasture as the experimental unit. Average daily gain and final weights were higher in year 1 than year 2 (P<0.01), and attributed to pasture burning before the start of grazing in year 1 but not in year 2. There were no interactions between treatments and no other year effects for any of the variables measured (P>0.1). No difference in mineral consumption between years (P=0.13) was noted. Average mineral and salt consumption across years was 147 and 59 g/day, respectively. Year 1 and 2 combined ADG for the mineral treatment animals was 0.14 kg higher than those in the salt treatment (P<0.01). Total weight gain improvement for the mineral vs. salt treatment was 11 and 19 kg, or 9 and 16%, for year 1 and 2, respectively. During year 1, animals treated for foot rot decreased from 38 to 12 by mineral supplementation (P=0.04). Foot rot treatments were not recorded in year 2. Providing a mineral supplement to steers grazing native spring-summer pastures in the Osage County, Oklahoma increased ADG and total weight gain that justified the cost of the mineral supplementation. Year 1 Year 2 (85 days grazing) (123 days grazing) Mineral Salt Mineral Salt SEM P-value Initial BW, kg 261 261 257 259 1.92 0.49 Final BW, kg 398 388 388 372 3.45 <0.01 ADG, kg 1.62 1.49 1.07 0.92 0.03 <0.01 Total Gain, kg 138 127 132 113 3.03 <0.01 61 138 57 Supplement Consumption, g/d 156 Key words: beef, grazing, mineral 214 6.74 <0.01 25 USE OF DRIED DISTILLERS GRAINS AS A SUPPLEMENTAL FEEDSTUFF FOR GROWING CATTLE ON WHEAT PASTURE. E. D. Sharman*, P. A. Lancaster, G. W. Horn, and J. T. Edwards, Oklahoma Agricultural Experiment Station, Stillwater, OK, USA. Two experiments were conducted to evaluate the efficacy of dried distillers grains and solubles (DDGS) as a supplemental feedstuff for stocker cattle grazing hard red winter wheat pasture (Triticum aestivum). In Exp. 1, 172 fall-weaned steer calves (BW = 255 ± 29.9 kg) were stratified by BW and assigned randomly within BW groups to 16, 7.3- to 9.7-ha, pastures in a randomized complete block design on November 15 at a stocking rate of 1.37 steer/ha for 113 d. Treatments were (1) non-supplemented, control (CON); (2) 1.13 kgsteer-1d-1 of whole corn (CORN); (3) 1.13 kgsteer-1d-1 of pelleted soybean hulls (SBH); and (4) 1.13 kgsteer-1d-1 of DDGS. All steers had free-choice access to a monensin-containing mineral (1764 mg monensin/kg). Data were analyzed by ANOVA using the Mixed procedure of SAS. Experimental units were pastures and sampling units were steers. Orthogonal contrasts included: (a) energy supplementation vs the CON; (b) type of supplement (i.e., high-starch, CORN, vs the average of the two high-fiber supplements, SBH and DDGS); and (c) SBH vs DDGS. Mean daily monensin intakes were 121, 76, 83 and 84 mg/steer for the CON, CORN, SBH, and DDGS treatments, respectively. Consumption of energy supplements averaged about 0.35% of mean BW for treatment 2-4. Energy supplementation tended to increase ADG (1.12, 1.11, 1.20, 1.20 ± 0.026 kg/d for treatments 1-4, respectively; P = 0.10), and was greater (P = 0.01) for SBH + DDGS vs CORN. There was no difference (P = 0.91) between SBH and DDGS. Supplement conversion, expressed as kg of as-fed supplement per kg of increased BW gain, was poor and averaged 11.9 and 13.2 for SBH and DDGS, respectively. Ultrasound measurement of intramuscular fat tended to be increased (3.55 vs 3.72%; P = 0.09) and backfat was increased (0.51 vs 0.57 cm; P = 0.03) by energy supplementation. In Exp. 2, 157 fall-weaned steer calves (BW = 242 ± 16.2 kg) were stratified by BW and assigned randomly within BW groups to 18, 7.3- to 9.7-ha, pastures in a randomized complete block design on December 4 at a stocking rate of 1.10 steer/ha for 50 d. Treatments were: (1) high-calcium, non-medicated free-choice mineral mix (control); (2) 0.91 kgsteer-1d1 of a monensin-containing energy supplement; and (3) 0.91 kgsteer-1d-1 of a monensin-containing supplement with 65% DDGS. Treatment 2 supplement contained (% as-fed): ground corn, 44.2; wheat middlings, 43.3; pellet binder, 5.0; salt, 1; dicalcium phosphate, 2.0; limestone, 3.7; magnesium oxide, 0.5; copper sulfate, 0.025; vitamin A to provide 29,964 IU/kg; and Rumensin 80® to provide 176 mg monensin/kg of supplement. Treatment 3 supplement contained 65% DDGS and 22.6% wheat middlings, while other ingredients were similar to treatment 2. Both energy supplements were pelleted in order to improve handling characteristics and minimize waste of the DDGS. Data were analyzed as described for Exp. 1. Orthogonal contrasts included the effect of energy supplementation and type of energy supplement. Mean mineral consumption was 0.18 ± 0.07 kgsteer-1d-1. Supplementation increased ADG (0.97, 1.12, and 1.20 ± 0.04 kg/d for treatments 1-3, respectively; P = 0.001) but was not influenced (P = 0.15) by substitution of DDGS into the corn/wheat middling supplement. Supplement conversions were improved compared with Exp. 1 and were 6.06 and 3.92 kg of supplement per kg of increased BW gain for the corn/wheat middling and DDGS-based supplements, respectively. Results of these experiments indicate that DDGS can be used as an alternative feedstuff in energy supplements fed at relatively low amounts to wheat pasture stocker cattle without compromising ADG or causing polioencephalomalacia. Ionophores are generally thought of as a technology for improving ADG of growing cattle, yet they also improve supplement conversion rates. The greatly improved supplement conversions observed in Exp. 2 accentuate the importance of including monensin in energy supplements for wheat pasture stocker cattle rather than providing energy supplements alone. The substantially improved supplement conversion for DDGS compared with corn/wheat middling supplement in Exp. 2 needs further evaluation. Key words: distillers grains, growing cattle, wheat pasture 215 26 A GENETIC MARKER FOR RESISTANCE TO FESCUE TOXICOSIS IN BEEF CATTLE. C.J. Kojima1, J.C. Waller*1, D.E. Spiers2, R.L. Kallenbach2, B.T. Campbell1, T.A. Cooper1, and J.K. Bryant2, 1University of Tennessee, Knoxville, TN, USA, 2University of Missouri, Columbis, MO, USA. Tall fescue is the prevalent grass in the Southeastern United States and can harbor an endophytic fungus (N. coenophialum) which causes symptoms known collectively as fescue toxicosis (FT): reduced feed intake and weight gain, decreased fertility, vasoconstriction and decreased ability to thermoregulate. The hallmark signs of fescue toxicosis are decreased serum prolactin concentrations and a rough, overly thick hair coat. A single nucleotide polymorphism (SNP) was discovered within the dopamine receptor D2 gene (DRD2) where a guanine/adenine substitution existed, creating two alleles, G and A. Two experiments were performed to assess the informativeness of this SNP for resistance to fescue toxicosis. For both experiments, data were analyzed by mixed model analysis of variance of SAS with and without repeated measures as appropiate. Forty-two Angus-based steers were grazed on Kentucky 31 tall fescue containing a toxic form of the endophyte (endophyte-infected fescue or EIF; n = 21), or Jesup tall fescue containing a non-toxic variety of endophyte (Jesup MaxQ; n = 21). Genotypic frequencies for this population were 0.33 (AA), 0.45 (AG) and 0.22 (GG). While genotype had no effect on serum prolactin or hair coat score in Jesup MaxQ steers, homozygous GG steers grazing EIF had decreased serum prolactin concentrations in April and May compared to AA steers (P < 0.05). Effect of genotype was then assessed in 53 Angus-based steers grazed on EIF. Genotypic frequencies for this population were 0.38 (AA), 0.36 (AG) and 0.26 (GG). Genotype was associated with serum prolactin concentrations (P = 0.004) and hair coat score (P = 0.0102) such that GG animals had decreased prolactin and increased hair coat scores relative to AA and AG animals. As most herds in Tennessee are spring-calving and therefore more prone to the effects of FT, we hypothesized that non-intentional selection for the advantageous allele was occurring in spring-calving herds relative to fall-calving herds, herds that are not exposed to FT, and herds of heat-resistant breeds. Genotypic and allelic frequency was assessed in 92 Tennessee Angus bulls, 78 Missouri Fall-calving Angus cows, 12 Missouri Spring-calving Angus steers, 10 Oklahoma Angus steers and 11 Florida Romosinuano steers. These data suggest that the advantageous allele is less prevelant in herds of cattle which are less affected by FT (eg. fall-calving, not grazed on tall fescue, or heat-tolerant breeds) than are spring- calving Angus cattle grazed on tall fescue. This may be a by-product of selection for longevity, fertility, and growth in cow-calf herds affected by FT. The DRD2 SNP may have potential use in the selection of animals resistant to fescue toxicosis. Genopytic and allelic frequencies of cattle from different herds and calving seasons. Genotypic Alleilc Frequency Frequency Calving Herd Season n TN Angus Steers I Spring TN Steers II AA AG GG A G 42 0.33 0.45 0.22 0.56 0.45 Spring 53 0.38 0.36 0.26 0.56 0.44 TN Bulls Spring 92 0.30 0.49 0.21 0.55 0.45 MO Angus Cows Fall 78 0.23 0.51 0.25 0.49 0.51 MO Angus Steers Spring 12 0.50 0.25 0.25 0.63 0.38 OK Angus Steers Unknown 10 0.20 0.40 0.40 0.40 0.60 FL Romosinuano Steers Unknown 11 0.33 0.33 0.33 0.50 0.50 Key words: cattle, dopamine receptor, tall fescue toxicosis 216 27 PERFORMANCE OF CATTLE GRAZING SWITCHGRASS OR COMBINATIONS OF BIG BLUESTEM AND INDIANGRASS DURING THE SUMMER IN THE MID-SOUTH. W.M. Backus, J.C. Waller*, P.D. Keyser, G.E. Bates, C.A. Harper, F.N. Schrick, and B.T. Campbell, University of Tennessee, Knoxville, TN, USA. The Tennessee Biofuels Initiative and others have selected switchgrass, a native warm-season perennial grass, as the main feedstock for cellulosic ethanol production. To meet the proposed level of ethanol production thousands of acres of pastures in Tennessee will be replaced with switchgrass. If beef producers could use the switchgrass as summer forage and later harvest it as a biofuel crop, the impact of this shift in grazing land would be reduced. Cattle grazing switchgrass or other warm-season native grasses should have greater weight gain than those grazing tall fescue during the summer. Tall fescue is a cool-season forage that has limited growth and reduced quality in the summer. Animal performance response to grazing switchgrass in this region is very limited. The objectives of this study are to graze switchgrass and other warm-season grasses that have the potential for cellulosic ethanol production to determine animal performance during summer. Two grazing studies were conducted in 2009 to compare animal performance of cattle grazing pastures of switchgrass (Panicum virgatum L.) or a combination of big bluestem (Andropogon gerardii Vitman) and indiangrass (Sorghastrum nutans L. Nash). Study 1 was conducted from May 29 to August 3 at the Highland Rim Research and Education Center near Springfield TN in which Angus and Angus cross steers (269±5.7 kg) were used in a randomized block design with two forage treatments: 1) Switchgrass (SG) or 2) a combination of big bluestem and indiangrass (BB/IG). All pastures used were two-yr old stands. Prior to initiation of grazing all steers were fed a high roughage filler diet for 4 d with individual BW taken in the early AM each day. The average BW for those 4 d was used for initial BW for the grazing period. Intake of the filler diet on a dry matter basis was about 2.5% BW. At the end of the grazing period all steers were fed the same filler diet for 4 d and those weights were used as the final weights for the study. Four steers (testers) were allotted to 1.2-ha paddocks with five replications per treatment. Additional steers were used in a put-and-take manner to keep forage in a vegetative state. Steers had free choice access to pasture, water, mineral, and shade. Steers were weighed on 21-d intervals to determine ADG. Data were analyzed using the MIXED procedure of SAS. Least square means for ADG of steers were not different (P > 0.05). The ADG of steers grazing SG and BB/IG, was 0.95 and 1.06 kg/d, respectively. Study 2 was conducted from June 5 to August 10 at the Middle Tennessee Research and Education Center near Spring Hill, TN in which bred Holstein replacement heifers (485±15.0 kg) were used in a randomized block design with four forage treatments: 1) switchgrass (SG), 2) a combination of big bluestem and indiangrass (BB/IG), 3) switchgrass plus red clover (Trifolium pratense L.) and 4) a combination of big bluestem and indiangrass plus red clover. The red clover was drilled in March at a rate of 5.6 kg/ha. All pastures used were two-yr old stands. Initial and final BW of heifers was established using the same filler diet and technique used in Study 1. Three heifers (testers) were allotted to 1.2-ha paddocks with four replications per treatment. Additional heifers were used in a put-and-take manner to keep forage in a vegetative state. Heifers had free choice access to pasture, water, mineral, and shade. Heifers were weighed on 21-d intervals to determine ADG. Data were analyzed using the MIXED procedure of SAS. Least square means for ADG of heifers were different (P < 0.05) when grazing SG or BB/IG without red clover. The ADG of heifers grazing SG and BB/IG, was 0.75 and 0.97 kg/d, respectively. Least square means for ADG of heifers were different (P < 0.05) when red clover was present stands of SG and BB/IG. Heifer ADG was 0.80 and 0.99 kg/d for SG + red clover and BB/IG + red clover, respectively. The results of these two grazing studies demonstrate the ability of native warm season grasses to provide suitable summer forage and animal performance for beef steers and replacement heifers within the tall fescue growing region. Key words: switchgrass, big bluestem, indiangrass 217 28 BOTANICAL COMPOSITION, NUTRITIONAL VALUE, VOLUNTARY FEED INTAKE ESTIMATION AND BEHAVIORAL HABITS OF CATTLE GRAZING ON RANGE PASTURES. D. Rodriguez-Tenorio*1, R. Gutierrez-Luna1,2, R.D. Valdez-Cepeda1, F.G. Echavarria1,2, M.A. Salas1, J.I. AguileraSoto1, M.A. Lopez-Carlos1, C.F. Arechiga1, and J.M. Silva-Ramos1, 1Universidad Autonoma de Zacatecas, Zacatecas, Zac., Mexico., 2Instituto Nacional de Investigaciones Forestales, Agricolas y Pecuarias, Morelos, Zac., Mexico. The objective was to compare botanical composition, nutritional value, voluntary intake and behavioral habits of cattle (Stobs, 1973; Chacon et al., 1976); under two different grazing systems in short grasslands of Zacatecas State, Mexico. Experiments were carried out at two ranches: R1) with a short-duration grazing system and R2) with a continuous grazing system. Samples were collected during the summer, autumn and winter for two consecutive years. Data collected for nutritional value, botanical composition and voluntary intake were analyzed using principal components analysis. Log-linear models for categorical variables were used to analyze grazing and behavioral habits. Ether extract content was affected by season. Crude fiber was similar in both ranches, but nutrient availability was reduced from 48% to 27%, and sometimes up to 62%. Nutrient availability was reduced from 45 to 27% at R1 as maturity advanced, whereas for R2, it was reduced from 46% to 25%. Grazing activity was different between ranches (P<0.05). Regarding behavioral habits in R1 cattle grazed 85 and 10% during winter and summer, respectively. Whereas, in R2 cattle grazed 20 and 70%. A greater grazing activity was performed at 10, 12 and 16 h for both seasons. There was a 50%-60% crude protein decrease during winter and early spring at both ranches. Results for NDF were higher than ADF at both ranches. Energy requirements for maintenance are 2.22 Mcal/kg for the growing season and 1.58 Mcal/kg for the period of dormancy. They are different to those found in our study: 1.00 Mcal/kg during the summer 0.90 Mcal/kg during the winter at both ranches, due to their poor range condition and the irregular rainfall. Key words: grazing systems, botanical composition, arid highlands 218 Grazing Livestock Nutrition Conference Index of Authors Acetoze, G., 205, 206 Clark, C., 191 Aguiar, A. D., 207 Cooper, T. A., 216 Aguilera-Soto, J. I., 218 Corl, B., 211 Anders, M., 213 Cruz, G., 205 Anderson, D. M., 57, 192 Cruz, G. D., 206 Arechiga, C. F., 218 DelCurto, T., 1 Backus, W. M., 217 Detweiler, C., 192 Bates, G. E., 217 Dittmar, R. O., 202 Bear, D. A., 209 Dobson, L., 202 Beck, P., 213 Doniec, M., 192 Biermacher, J. T., 204 Dove, H., 31, 133 Blanton, J. R., 204 Du, M., 210 Boland, H. T., 194 Echavarria, F. G., 218 Boland, H., 212 Edwards, J. T., 215 Brulc, J. M., 10 Endecott, R. L., 152 Bryant, J. K., 216 Forbes, T. D. A., 207 Bu, D., 199 Gillen, R. L., 208 Bu, D. P., 200, 201 Glassey, C., 191 Butler, W. R., 93 Gregorini, P., 191 Butler, T. J., 204 Gruber, M., 195 Campbell, B. T., 216, 217 Guiroy, P. J., 214 Carstens, G. E., 207 Gunter, S. A., 208 Caton, J. S., 104, 210 Guo, Y. Q., 201 Chase, L. E., 203 Guretzky, J. A., 204 219 Gutierrez-Luna, R., 218 McLennan, S. R., 133 Harper, C. A., 217 McLeod, K., 191 Hensen, A., 197 McMurry, B., 214 Hess, B. W., 104, 210 Medeiros, F. S., 193 Hess, T., 213 Meyer, A. M., 210 Hoffman, K., 203 Milliniks, J. T., 180 Horn, G. W., 215 Morgan, M., 213 Hu, T., 201 Morrical, D. G., 209 Hubbell, D., 213 Murphy, E. J., 196 Jago, J., 191 Nelson, K. E., 10 Jonker, A.., 195 Nolen, B., 192 Kallenbach, R. L., 216 Olson, K. C., 1 Kering, M. K., 204 Patino, H. O., 193 Kerley, M., 125 Petersen, M. K., 180 Keyser, P. D., 217 Poppi, D. P., 133 Kojima, C. J., 216 Randel, R. D., 207 Kronberg, S. L., 196 Reuter, R. R., 204 Lambert, B. D., 202 Rimbey, N. R., 170 Lancaster, P. A., 215 Roberts, A. J., 180 Lardy, G. P., 152 Rodriguez-Tenorio, D., 218 Lillie, A., 211 Rogers, J. K., 204 Liu, K. L., 201 Romera, A., 191 Lodge-Ivey, S. L., 24 Rossow, H., 205 Lopez-Carlos, M. A., 218 Rossow, H. A., 206 Maddock, R. J., 196 Rouquette, Jr., F. M., 207 McCuistion, K. C., 202 Rus, D., 192 220 Russell, J. R., 209 van den Pol-van Dasselaar, A., 197, 198 Salas, M. A., 218 Waller, J. C., 216, 217 Scaglia, G., 194, 211, 212 Wang, Y., 195 Scholljegerdes, E. J., 196 Wang, J., 199 Schrick, F. N., 217 Wang, J. Q., 200, 201 Sharman, E. D., 215 Waterman, R. C., 93, 180 Shen, J., 199 Watkins, B., 213 Showers, S. E., 214 Wei, H., 199 Silva-Ramos, J. M., 218 Wei, H. Y., 200 Soder, K. J., 203 White, B. A., 10 Spiers, D. E., 216 Wickersham, T. A., 202 Springer, T. L., 208 Wiley, L., 202 Sun, P., 199, 200 Winters, C., 192 Swecker, Jr., W. S., 211 Yan, Z. H., 200 Tedeschi, L. O., 207 Yeoman, C. J., 10 Thacker, E., 208 Yu, P., 195 Torell, L. A., 170 Zhang, Y. D., 201 Valdez-Cepeda, R. D., 218 Zhou, L. Y., 200 221 Untitled-1 1 6/5/2003, 2:42 PM Untitled-1 1 6/5/2003, 2:42 PM Conference Sponsors Western Section American Society of Animal Science USDA – National Institute of Food and Agriculture Alpharma Animal Health Cargill Kahne Animal Health Agricultural Experiment Stations of Arizona, Colorado, Montana, and Wyoming