.. ti: ..)....2 13:4.) 515.2... An; . E 2333...: Irfluzw .5 £533.43. :1... 523.5 .53. .1 :L... 5 .3. ~ 59.6. .mrdflflu 3.7... z.\.|..x 3.5 .. 3 5.12? 3 31‘: 3.1,.» : V3.4... . . .1 . 2:; .. 1: 1:: .u ‘. 5..., 5:. 12:. I. 5. 7.113.: {EL 1.1.»: 73...? 2:1. {7:22.35 K. : ME, 3.. SITY LIBRARI lull lllllllll i 072 ? Illllllillllllllllllll 3 1293 0101 This is to certify that the dissertation entitled IN VITRO METHODS TO DETERMINE PROTEIN DEGRADATION OF FEEDS FOR RUMINANTS presented by Richard A. Kohn has been accepted towards fulfillment of the requirements for Ph.D. degreein Animal Science Wfi/ézp flajor professor Date 8/18/93 MSU is an Affirmative A ction/Equal Opportunity Inslilution 0-12771 LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before due due. DATE DUE DATE DUE DATE DUE IN VITRO METHODS TO DETERMINE PROTEIN DEGRADATION OF FEEDS FOR RUMINANTS By Richard A. Kohn A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Animal Science 1993 ABSTRACT IN VITRO METHODS TO DETERMINE PROTEIN DEGRADATION OF FEEDS FOR RUMINANTS By Richard A. Kohn Three laboratory methods for evaluation of feed protein degradation were studied: 1) in vitro fermentation with 35S used to label microbial protein, 2) in vitro incubation with cmde extract of enzymes from rumen fluid, and 3) sequential extraction of forage proteins into 6 fractions postulated to degrade at different, yet uniform first order rates. In the first method, feeds were incubated in batch culture in bicarbonate-phosphate media with 35S-cysteine and rumen inocula. At the end of incubation, feed proteins were separated from free bacteria by sequential centrifugation. Ratio of N to 35$ in free bacteria was multiplied by 358 in other fractions to estimate microbial N. Two rinses in HCl solution containing free cysteine, NazSO4 and NaSO3 were required to remove most non-protein 358. Ratio of N to 35S varied across treatments and lengths of incubation because feed S was incorporated into microbes. The proposed method was simpler than previous work, but high variation within treatments persisted. For the second method, microbial enzymes were dislodged by extraction of rumen solids with butanol and acetone, and subsequent solubilization of proteases in water. This extract was incubated with feeds to degrade proteins. The enzyme extract maintained 62% of the activity in whole rumen fluid, but was more concentrated. Proteolytic activity appeared to be resistant to self-degradation, but rates of degradation of alfalfa hay and soybean meal declined with length of incubation. A third method evaluated N solubility for determination of crude protein degradation. Several forages were degraded by crude enzyme extract and were subjected to sequential solubilization in: l) trichloroacetic acid, 2) buffer, 3) acetone, 4) detergent at pH 7, and 5) detergent in IN H2804. Additionally, 8 forages were degraded in rumen extract followed by sequential fractionation. Protein degradation of individual fractions was determined by multiple regression analysis of degradation against fractions, or by direct measurement. These studies demonstrated that buffer soluble proteins were readily and completely degraded by rumen extract, but other fractions were degraded incompletely and inconsistently over time. DEDICATION He [Professor Jefi‘rey Williams ] wanted to publish her research ndings before she [Graduate Student Maie El ’Kassby] had a chance to, so he fire her. Professors do that all the time. They ’re parasites that live on the backs of their graduate students. --Zolton Ferency, 1922-1993 This dissertation is dedicated to the memory of civil rights activist Zolton Ferency, a man who taught us by example to be brave, to say and to do what we know is right, to struggle, and never to compromise our principles. Zolton, as a professor, lawyer and politician often found himself involved with a corrupt system, but he never became corrupted by the system. projr was i degrc O'Nc new; with 1 every Stalin Sniff; r3363: diet 5%; grad, ACKNOWLEDGMENTS I would like to thank my wife Tammy for her support of me regarding this project, but especially for her support regarding every other aspect of my life. Tammy was by far the most significant ( P < .0001) discovery I made while working toward my degree. She reminds me daily of what is truly important. I would like to express my appreciation to my fellow graduate students, Kitty O'Neil, Katharine Knowlton, Rick Dado, Steve Mooney and Mary Beth Roe. I could never begin to describe how much I learned from them. I also would like to thank Dave Main and Dewey Lon guski for their assistance with maintaining and optimizing the lab, help on experiments, and for teaching me everything I'd ever want to know about country music. I thank Jim Liesman for sharing his broad knowledge of laboratory methods, statistics, and animal science, and for the interesting philosophical conversations. I thank my graduate committee members: Dr. Roy Emery, Dr. Kenneth Nadler, Dr. Steven Rust, and Dr. Michael VandeHaar, and previous members Dr. Charles Sniffen and Dr. Derek Lamport. All of these members have helped me in different Ways to improve the quality of my education, and to broaden my understanding of the research. I would especially like to thank Dr. Mike Allen, my major professor and dissertation director, for his interest in my education and research. Last but not least, I would like to thank the woman with the ruler at the graduate School , who Will determine in two minutes that this dissertation is acceptable so I can graduate. I wish only that my committee members were as efficient. rm CH4 CHM TABLE OF CONTENTS INTRODUCTION 1 CHAPTER 1 LITERATURE REVIEW 4 Importance of Ruminal Protein Degradation 4 Protein Metabolism 4 Protein Digestion in Ruminants 5 Ruminant Protein Requirements 5 Effect of Protein Source on Degradation 11 Animal and Microbial Effects on Protein Degradation ............................. 18 Methods to Determine Protein Degradation 21 Measurement of Protein 21 Estimation of Microbial Protein Synthesis 22 In Vivo Methods to Measure Protein Degradation .............................. 28 In Situ Methods to Measure Protein DPgmdafinn 29 In Vitro Methods to Measure Protein Degradation 31 LITERATURE CITED 42 IHAPT ER 2 Storage of fresh and ensiled forages by freezing affects fibre and crude protein fractions 57 Abstract 57 1. Introduction 58 vi Cf. 2. Experimental 2.1. Treatment of samples 2.2. Analyses 2.3. Statistical 2.4. Freeze—drying 3. Results and D' ‘ 3.1. Crude protein fractinne 3.2. Ash, fibre and lignin 3.4. Freeze-drying 3.6. Relevance of research 4. References IHAPTER 3 Observations on determination of protein degradation of feeds by in vitro batch fermentation with 3SSulfur used to label microbial growth. ............ ABSTRACT INTRODUCTION METHODS Final Method for Microbial Fermentation Development and Validation RESULTS AND DISCUSSION Separation of Incorporated from Non-Incorporated 358 ..................... Addition of reducing agents prior to rinsing. .............................. Precipitation of cysteine with protein ......................................... Rinsing the pellet of 3SS—cysteine precipitate Solubilization of Incorporated 353 Tissue solubilizer and HCl digestion .......................................... HCl digestion and chromatography with Dowex ........................ vii 59 59 59 61 61 61 61 62 63 65 66 71 71 72 76 76 78 79 79 79 80 82 92 92 92 CH.A CH4. Digestion in sulfuric acid Direct Measurement of 355 in Microbes ............................... Ratio of N to 358 in Labeled Microbes REFERENCES CHAPTER 4 Enrichment of proteolytic activity in preparations from the rumen for in vitro studies Synopsis Introduction Materials and ‘ ' ‘L ‘ Validation and improvement of azocascin method Azocasein degradation Chilling, blending and timing Incubation in vitro Extraction in buffer, detergent or amtnne Results and 1" Validation and improvement of azocascin method ........................ Chilling, blending and rinsing Incubation in vitro Extraction in buffer. detergent or Mm-.. REFERENCFQ CHAPTER 5 An in vitro method to measure protein degradation of feeds using enzymes extracted from rumen ‘ ‘ Synopsis 'L J a' Materials and ‘ ' ‘L ‘ 133 133 133 135 Solutions and feeds 135 Butanol-acetone extraction 135 Feed protein ‘ 9 ‘ " 135 Reducing agents 138 Cellulase and amylase 138 Proteolytic activity over time 139 Effect of enzyme level 140 Comparison of feeds and variation 140 Results and I" 141 Butanol-acetone extraction 141 Reducing agents 142 Proteolytic activity over time 143 Cellulase and amylase 143 Effect of enzyme level 144 Comparison of feeds and variation 144 Conclusion 147 REFERENCFQ 148 CHAPTER 6 Effect of Plant Maturity and Preservation Method on In Vitro Protein Degradation of Forages 158 ABSTRACT 158 INTRODUCTION 159 METHODS 159 RESULTS AND DISCUSSION 161 REFERENCES 164 CHAPTER 7 Prediction of Protein Degradation of Forages from Solubility Fractions ........... 178 ix CHAPTER PRE ABSTRACT 178 INTRODUCTION 179 METHODS 180 Forages and Protein Degradation 180 Protein Fractinnatinn 181 Statistics for the First Experiment 133 Solubility Fractions After Degradation 184 RESULTS AND DISCUSSION 185 Fractionation of Forages 185 Prediction of Degradation from Fractions 187 Solubility Fractions After Degradation 189 Comparison to Previous Results 192 Implicatinnc 193 REFERENCES 195 BR 8 EPILOGUE 214 IX PRELIMINARY EXPERIMENTS AND OBSERVATIONS ............................ 219 Observations on Protein Solubility 219 Measurement of Forage CP Solubility 219 Effect of pH and Eh on CP Solubility 220 Effect of Temperature on Proteolytic Activity of Alfalfa .................... 221 Observations on In Vitro Fermentation 222 Titration of in vitro media 222 Variation in pH and reducing potential for in vitro batch (‘nltan 223 Effect of Reducing Agents on Microbial Fermentation 224 x Effect of carbohydrate source on microbial growth and 358 N ratio. Observations on Crude Rumen Fluid Extract Reducing agents on protein degradation. Centrifugation of rumen extract before use. ........................................ Presoaking of alfalfa. Tungstic acid vs. TCA precipitation xi 225 226 226 229 230 230 CHrl CHAI LIST OF TABLES {AFTER 2 TABLE 1. Effect of freezing on crude protein and fibre fractions of smooth bromegrass (g 100 g—l M). TABLE 2. Effect of freezing on crude protein and fibre fractions of fresh luceme (g 100 g-l DM). 68 69 TABLE 3. Effect of freezing on crude protein and fibre fractions of luceme silage (g 100 g-l 7O {AFTER 3 Table 1. Precipitation of 358 (cpm) with trichloroacetic acid after incubation of Trypticase with microbial inocula for 10 h with addition of sulfide or sulfide and acid before rinsing the pellet in water. ........................ Table 2. Precipitation of 358 (cpm) with trichloroacetic acid after incubation of soybean meal with sterile media for 4 h with addition of reducing agents before rinsing the pellet in water three times. ....................... Table 3. Precipitation of 35S—cysteine (cpm) with different methods to Table 4. Method of rinsing the trichloroacetic acid pellet on removal of 3SS—cysteine (cpm) from a feed mixture (starch, cellulose and soybean meal) precipitate protein, with and in the absence of soybean meal.1 ...................... Table 5. Precipitation of 35S (dpm) from media incubated with rumen microorganisms for 8 h with Trypticase, or without organisms with soybean meal. Table 6. Effect of number of times of rinsing a feed sample on removal of 3SS-cysteine from TCA precipitable material. Table 7. Effect of soaking feed sample during rinsing on precipitation 3SS-cysteine. Table 8. 3SSocysteine precipitated with different feed samples incubated (38 °C) for 3 h without microbial inocula ..... 81 81 83 84 86 86 86 87 Table 9. Effect of rinse solution pH on precipitation of 358- -cysteine (cpm) after rinsing twice. 89 Table 10. Effect of adding eth,‘ " ' ‘ ‘ " acid, dithiolthreotol, or cytyl trimethylammonium bromide to rinse solution on 3SS-cysteine precipitation (cpm) with starch Table 11. Effects of timing on precipitation of 358- -cysteine (dpm) with feed samples 1n pH 6. 8 media. 89 93 Table 12. Effects of timing on precipitation of 35S-cysteine (dpm) with feed samples in 1% HCl. Table 13. Recovery of 35S (cpm) in samples digested in H2804 with peroxide and recovery per m1 NaOH required to neutralize the H2804 in the digest. Table 14. Digestion of 35S-cysteine in sulfuric acid and peroxide on recovery of radioactivity (cpm) Table 15. Solubilization of 358 from radio—labeled microbes using different solvents. Table 16. Effect of incubation time with microbial inocula and nitrogen source on nitrogen and 358 activity insoluble in trichloroacetic acid. ............... Table 17. Effect of amount and source of nitrogen on nitrogen and 35S activity insoluble in trichloroacetic acid after incubation with microbial inocula for 18 h. Table 18. Estimation of microbial nitrogen and undegraded feed protein after incubation for 20 h with rumen inocula. AFTER 4 Table 1. Carbohydrate composition of media used for incubation of ruminal organisms in continuous culture. 93 98 98 101 102 104 126 Table 2. Treatment of rumen contents on proteolytic activity per m1 rumen fluid (PA) and per mg nitrogen (PA/N). 127 Table 3. Mean proteolytic activity per ml rumen fluid (PA) and per mg N 1n rumen fluid (P A/N). Table 4. Effect of treatment of rumen contents (chilling, rinsing and blending) on recovery of proteolytic activity per ml rumen fluid (PA) and per mg N (PA/N) in fractions separated by sequential centrifugation. 128 129 Table 5. Mean proteolytic activity per m1 rumen fluid (PA) and per mg N in rumen fluid (PA/N) by centrifugation fraction. 130 xiii Table 6. Effect of preincubation in continuous culture at .07/h on proteolytic activity per ml rumen fluid (PA) and per mg N in rumen fluid 131 (PA/N) by centrifugation fraction. Table 7. Effect of solubilizing rumen fluid precipitate in reduced media, media with detergent, and acetone on proteolytic activity per ml rumen fluid (PA) and per mg N (PA/N). HAPTER 5 Table 1. Effect of reduction of in vitro media on soybean meal (SBM) degradation at 16 h in enzyme extracted from the rumen. ............................... Table 2. Solubility of nitrogen from luceme hay in trichloroacetic acid as affected by incubation with protease extracted from the rumen (P), commercial cellulase (C), soaking in pH 4.7 buffer before P, and C before P. Table 3. Solubility of nitrogen from luceme hay in buffer (pH 6.8) as affected by incubation with protease extracted from the rumen (P), commercial cellulase (C), soaking in pH 4.7 buffer before P, and C before P. Table 4. Solubility of dry matter (DM) from luceme hay in buffer (pH 6.8) as affected by incubation with protease extracted from the rumen (P), commercial cellulase (C), soaking in pH 4.7 buffer before P, and C before P. Table 5. Effects of enzyme level and addition of enzyme after 8 h on soybean meal degradation at 8 and 16 h. Table 6. Average crude protein degradation (percentage of original) after 0, 6 or 16 h incubation with enzyme extracted from the rumen. ............. [AFTER 6 TABLE 1. Mean forage composition by species and maturity level ) 132 152 153 154 155 156 157 . 167 across forage preservation methods (11 = 12 . TABLE 2. Mean protein degradation at O, 2 and 24 h by forage species ) and maturity level across forage preservation methods (n :12 . .................... TABLE 3. Mean protein degradation at 0, 2 and 24 h by forage species and preservation method across maturities (n = 9) 168 169 APTER 7 TABLE 1. Mean percentage protein in each of 6 solubility fractions by forage species and maturity level across forage preservation methods (11 -12). .. 198 TABLE 2. Mean protein percentage in each of 6 solubility fractions by forage species and preservation method across maturities (n = 9). .................... 199 TABLE 3. Prediction of forage protein degradation at 2 h determined from multiple regression analysis of protein“ 200 TABLE 4. Prediction of forage protein degradation at 24 h determined from multiple regression analysis of protein“ 201 TABLE 5. Percentage decrease from the initial amount of crude protein in solubility fractions after 2 h degradation in vitro 202 TABLE 6. Percentage decrease from the initial amount of crude protein in the respective solubility fractions over time across forages. .......................... 203 TABLE 7. Estimation using the Lucas test of percentage protein degraded from fractions. 204 APPENDIX APPENDIX TABLE 1. Effect of straining through cheese cloth. filtering through Whatrnan # 54, or centrifuging (1500 g) on water soluble CP, or trichloroacetic acid (T CA) soluble CP in orchard grass hay. 220 APPENDIX TABLE 2. Effect of soaking one tenth bloom freeze-dried alfalfa on protein solubilized in TCA. 222 APPENDIX TABLE 3. Effect of starch and cellulose level (E) and casein level (P) on pH and Eh of in vitro batch culture. ..................................... 223 APPENDIX TABLE 4. Effect of sodium ascorbate addition on precipitation of 355 (dpm) with soybean meal fermented with rumen microorganisms. 224 APPENDD( TABLE 5. Precipitation of 358 after incubation of 358- cysteine with rumen microorganisms, trypticase and the carbohydrate sources listed as treatments. 226 APPENDIX TABLE 6. Method of reduction on soybean meal (SBM) degradation with enzyme extracted from the rumen 227 APPENDIX TABLE 7. Effect of the presence of enzyme and reducing method on protein degradation of alfalfa. 228 APPENDIX TABLE 8. Precipitation of CP before and after 16 h incubation with crude rumen extract using trichloroacetic acid (TCA) or tungstic acid (W). 232 APPENDIX TABLE 9. Interference of crude rumen extract with measurement of precipitable protein (mg N of total 8 mg) 1n corn or soybean meal (SBM) before ‘ 233 XV CHAP CHAP LIST OF FIGURES CHAPTER 3 Figure 1. Sulfur Metabolism in the Rumen 75 CHAPTER 4 Fig. 1. Optical density of azocascin solution, centrifuged rumen fluid pellet, and centrifuged rumen fluid supernatant at varying wavelengths of light. Fig. 2. Treatment of rumen fluid on proteolytic activity per ml rumen ............ Fig. 3. Treatment of rumen fluid on proteolytic activity per mg nitrogen in enzyme preparation. CHAPTER 5 Figure 1. Effect of incubating rumen enzyme preparation at 38°C on proteolytic activity on azocascin. Figure 2. Degradation of soybean meal and luceme hay with enzyme extracted from rumen fluid. CHAPTER 6 Figure 1. Percentage of alfalfa forage protein degraded in early mid and late maturity forage. Figure 2. Percentage of bromegrass forage protein degraded in early mid and late maturity forage. Figure 3. Percentage of reed canarygrass forage protein degraded in early mid and late maturity forage. Figure 4. Percentage of whole plant corn protein degraded in early and late maturity forage. CHAPTER 7 Figure 1. Scheme for sequential extraction of forage crude protein. ................ 123 125 150 151 170 172 174 . 205 Figure 2. Prediction of forage protein degradation at 24 h from solubility fractions of original material. ' 206 Figure 3. Change in crude protein solubility fractions due to degradation in ruminal enzyme extract. 208 Figure 4. Lucas test for uniformity of degradation in ruminal enzymes across forages. 210 Figure 5. Lucas test for uniformity of degradation in ruminal enzymes across forages. 212 met] rau'c indir film:- COmI INTRODUCTION In ruminant species optimal protein utilization results when adequate nitrogen is available to rumen microbes at a utilizable rate, while additional protein remains undegraded until it is digested and absorbed in the small intestine. Recently introduced systems to balance cattle rations have emphasized the need to consider ruminal protein degradation. Though many feed protein sources have been evaluated for protein degradation, the forage protein, which comprises a major portion of the diet, is highly variable and the degradation properties are difficult to predict. Therefore, a routine method to predict forage protein degradation in the rumen is needed to balance cattle rations for ruminally degraded and undegraded protein. Rumen degradability of feed protein depends on the rate of degradation of individual proteins in the rumen, and on the rate of passage of those proteins from the rumen. Both of these factors may be influenced by protein associations with other feed components, and by availability of chemical bonds of the proteins to microbial attack. Separation of feed proteins based on solubility may have practical significance to rumen protein degradation, because solubility in different media can separate proteins based on their associations with plant organelles (i.e. cytoplasm, membranes, cell wall), and solubility may relate to aspects of protein structure which affect degradation. The proposed hypothesis is that a sequential extraction of crude protein in forages may provide for protein fractions that degrade at uniform rates across forages. This information would enable us to routinely predict protein degradation from size of individual crude protein fractions. solul mme evalu crude 0n 50] their c metho metho re‘lUin occur I immec The objectives of this work were: 1) to develop methods to separate nitrogen fractions of forages based on solubility characteristics that may be related to ruminal protein degradation, 2) to develop a laboratory procedure to simulate feed protein degradation in the rumen, and 3) to evaluate ruminal degradation of the isolated nitrogen fractions and to evaluate the feasibility of using patterns of nitrogen partitioning to estimate overall crude protein degradation. The approach to the first objective of isolating protein fractions in plants based on solubility was based on previous research on types of proteins found in plants and their cellular locations as described in Chapter 1. Some preliminary studies to develop methods to separate and measure these proteins are described in the Appendix. Any method to routinely predict forage protein degradation based on solubility would require that forages be preserved before analysis in such a way that changes do not occur before processing. Since changes in forages during preservation would immediately result in an artifact for all subsequent measurements, research into how forages should be preserved before analysis was undertaken as described in Chapter 2. The second objective was to develop a method to measure protein degradation in an environment that simulates the rumen, but with the ability to monitor degradation of individual protein fractions isolated according to the first objective. The first attempt was to use an in vitro fermentation method that was developed as described in Chapter 3. This method was too labor intensive and variation within treatments was too high to use the method for the desired purposes. The results are summarized in this dissertation so that further work might improve development of the method for other purposes. A second approach to measure protein degradation was to prepare a crude extract from rumen fluid that would have high proteolytic activity for use with in vitro Slur 11511 {13C [in degrad am the analys-L after de how [he needed r. 3 studies. In order to measure proteolytic activity in rumen fluid preparations, a method using azocasein was refined and briefly validated for use with rumen fluid preparations as is described in Chapter 4. This method was then applied to determine gross methods to concentrate proteolytic activity. The results of these studies, as described in Chapter 4, led to development of a method to extract rumen fluid solids with acetone and butanol to free the proteases from microbial cells, and then to extract the enzyme from the acetone-insoluble fraction with water. The final method used was improved as described in the Appendix and in Chapter 5, and initial validation and studies on uses and limitations of the method are described in Chapter 5. The third objective was to measure protein degradation of individual CP fractions. A wide array of different forages were collected and evaluated for protein degradation as described in Chapter 6. Cnide protein was fractionated in these forages and the protein degradation of the fractions was determined from multiple regression analysis of fractions vs. degradation, and from solubility analysis of the residue left after degradation of some forages. These results are described in Chapter 7. Finally, the studies are interrelated in Chapter 8, with particular emphasis on how the later studies offer insight into the previous ones, and the further research needed for further development of these methods CHAPTER 1: LITERATURE REVIEW Importance of Ruminal Protein Degradation Protein Metabolism. Proteins are composed of amino acids which have a chiral carbon atom with a carboxylic acid group (COOH), an amine group (NH2), a hydrogen atom (H) and a variable residue (R). The residue differs for each of 20 common amino acids. Amino acids that can be synthesized by the body from common precursors are "non-essential", and amino acids that cannot be synthesized by the body from common precursors are termed "essential". "Limiting" amino acids are those which are absorbed from a given diet in the least concentration relative to requirements. That is, a decrease in the protein concentration of the diet would result in a decrease in production or growth due to the limiting amount of these amino acids. Proteins synthesized by body tissues have specific amino acid sequences. The non-essential amino acids can be synthesized de novo for the production of these proteins, but essential amino acids or their precursors must be supplied in the diet. If the body lacks an essential amino acid, protein synthesis will be arrested until that amino acid is supplied regardless of how much total protein is provided in the diet. The amine group of amino acids can be replaced with an oxygen atom that shares two electrons (double bond) by deamination, thus leaving a keto—acid. The keto— acids are metabolized for energy. Alternatively, the amine group may be transferred from an amino acid to a keto-acid by a process called transamination. In animals, glutamate and aspartate may be synthesized de novo from NH3 and the respective carbohydrate backbone, and these may be converted to glutarnine and asparagine non be 11 feed C005 ahso mien abort. amin. 3mm: aDim; this p inrpo; diggs Wail; {01' p( blow in W being is beli 5 respectively by amination (Lehninger, 1975). Once a keto-acid is aminated to form an amino acid, the amine may be transferred to other keto-acids to form the non-essential amino acids. Proteins and amino acids are continually synthesized and degraded during normal metabolism. Some essential amino acids are lost in this metabolism, and must be replaced by the diet. Protein Digestion in Ruminants. Ruminants have a unique way of digesting feed protein. Some feed protein is digested by rumen microorganisms, hydrolyzed to amino acids and eventually to ammonia. This non-protein nitrogen (NPN) is either consumed by the microorganisms and used to synthesize microbial proteins, or it is absorbed from the gut as NH3. Some feed protein is not degraded by ruminal microorganisms. This undegraded feed protein and the microbial protein pass to the abomasum and small intestine and can be digested by animal enzymes. Peptides and amino acids may be absorbed from the small intestine and used to meet the animal's amino acid requirements. Some of this protein is not digested and passes out of the animal in the feces. Some microbial protein synthesis occurs in the large intestine, but this protein is not available for absorption. However, microbial growth may be important for digestion of fiber in the large intestine, and some end-products from this digestion may be absorbed in the large intestine and used for energy. Therefore, availability of nitrogen may be important to digestion in the large intestine, especially for post-gastric fermenters such as the horse. Nitrogen in the blood may be transported to the large intestine and used for microbial growth. Absorbed ammonia has one of four possible fates. On high protein diets, most blood ammonia is transported via the blood to the liver, converted to urea, and excreted in urine. Some ammonia is excreted directly from the blood by the kidney without first being converted to urea. Some urea and ammonia are recycled back to the gut by what is believed to be passive transport. Urea and ammonia diffuse into saliva that enters the 6 gut, and across the rumen and intestinal wall. Microbes are thus able to use this NPN for synthesis of microbial protein. Finally, ammonia may be used for synthesis of non- essential amino acids. Ruminant Protein Requirements. The protein requirements of ruminants are related to the digestive process. As with non-ruminants, the animal must absorb essential amino acids from the small intestine. However, ruminal microorganisms are capable of synthesizing all of the common amino acids de novo. Therefore, ruminants have the unique ability to convert NPN to protein, and may survive and produce without consumption of true protein. None-the—less, microbial protein synthesis alone is not adequate to maintain the high levels of production of modern ruminants (Orskov et al., 1980; Schwab et al., 1992). Therefore, some feed protein that escapes ruminal degradation is required in the diet. This protein is referred to as "undegraded intake protein" or "UIP" by a system commonly used to balance dairy cattle diets in the United States (National Research Council, 1989). The protein which is degraded is referred to as "degraded intake protein" or "DIP". Adding sources of protein that are slowly degraded in the rumen has been shown to increase milk production in some cases (Kung and Huber, 1983; Ciszuk and Lindberg, 1988; Voss et al., 1988; Hamilton et al., 1992), but not in other cases (Folman et al., 1981; Robinson and Kennelly, 1988). A requirement of UIP also has been demonstrated for maximal growth of beef steers (Lindberg and Olsson, 1983; Hunter and Siebert, 1987 ; Newbold et al., 1987). The amount of amino-acid protein absorbed in the small intestine was greater for cattle fed diets containing a greater portion of UIP relative to DIP. In these diets the rumen ammonia levels were lower for the high UIP diets. In addition, infusion of amino acid protein directly to the small intestine also increases production (Schwab et al., 1992) and dry matter intake (Me pron optir pron intes IHad. '______~._—. k 7 (Meissner and Todtenhofer, 1989), thus demonstrating the importance of undegraded protein entering the small intestine. Though providing for protein that escapes ruminal degradation is important to optimizing production of high-producing ruminants, feeding to increase microbial protein synthesis may also increase the flow of essential amino acids to the small intestine. Microbial growth requires an adequate supply of carbohydrate that is fermented in the rumen, and a rumen environment conducive to microbial growth. Microbial growth may be limited if carbohydrate degradation is too low to supply the energy for growth (National Research Council, 1985). Alternatively, on high starch diets, the pH of the rumen fluid will decrease which will lower microbial growth (Staples and Lough, 1989). The interactions of these factors result in a limit to the maximal amount of protein synthesis that can occur. Further research will address ways to optimize microbial protein synthesis, but additional protein requirements must be met by provided UIP. Often times, microbial protein synthesis is limited by the amount and form of nitrogen available to rumen organisms (Polan, 1988; Rooke and Armstrong, 1989). Inadequate DIP may result in a decrease in protein entering the small intestine due to a depression of microbial protein synthesis (Rooke and Armstrong, 1989). If inadequate N is available to rumen microbes, fiber digestion may also be reduced (McAllan and Smith, 1983B; McAllan and Griffith, 1987; Adamu et al., 1989; Zorrilla—Rios and Horn, 1989). When microbes are not available to digest fiber in forages, fiber has a greater filling effect, and hence feed intake is reduced (Hunter and Siebert, 1987; Ciszuk and Lindberg, 1988). Therefore, while it is important to supply an adequate amount of UIP to help meet the needs of the host animal, there is also a requirement for DIP to provide for microbial growth to synthesize microbial protein, degrade fiber in the diet, and promote normal rumen function. acid: 1985 short mixtt amin. not re infusi synth. Anus; 0r fist. fibcrc' 8 While most of the protein needs of rumen microorganisms can be met with ammonia and carbohydrates, microbial growth is maximized by the inclusion of amino acids or peptides in the rumen (Mathison and Milligan, 1971; Thomsen, 1985; Polan, 1988). In vitro studies involving incubation of rumen fluid with amino acid mixtures showed that maximal microbial growth was obtained by supplementation of urea with a mixture of leucine, methionine and histidine (Fujimaki et al., 1992). Addition of other amino acids did not result in further increases in microbial growth. Sulfide addition did not result in as high a microbial growth as either cysteine or methionine. Ruminal infusion of casein or supplementation with soybean meal increased microbial protein synthesis in the rumen above that observed from infusion of urea (Rooke and Armstrong, 1989). Supplementation with soybean meal (McAllan and Griffith, 1987) or fishmeal (McAllan and Smith, 1983; McAllan and Griffith, 1987) resulted in greater fiber digestion than supplementation with urea. Optimal formulation of ruminant diets requires attention to protein and energy interactions. As was mentioned, the amount of microbial protein synthesis depends in part on niminally-available carbohydrate to support microbial growth. Therefore, the amount of DIP required in a ration depends on the level of ruminally-available carbohydrate. Two carbohydrate sources were compared for effect of energy level on protein flow to the small intestine (Cecava et al., 1988). A study showed faster disappearance of soybean meal and sunflower meal from nylon bags suspended in the rumen when sheep were fed dried grass instead of a barley-based diet with little fiber (Ganev et al., 1979). This suggests that protein degradation may be affected by carbohydrate sources in the ration. Protein requirements for ruminants also are affected by post-absorptive metabolism of carbohydrates and protein. Some required protein is used for gluconeogenesis (synthesis of glucose from pyruvate). Because much of the glucose con. 5an gluc rrter intes glucr (Oke ratio: lactos fequir acids, consumed by ruminants is fermented to volatile fatty acids in the rumen, considerable synthesis of glucose is required, and amino acids may be used as precursors for gluconeogenesis. Intravenous infusion of glucose in lambs resulted in greater nitrogen retention (Matras and Preston, 1989). A greater amount of starch fermented in the intestine (due to rumen protection by formaldehyde treatment) resulted in greater blood glucose and insulin levels and improved efficiency of N utilization by Angus steers (Oke and Loerch, 1992). It has been suggested that higher undegraded protein in the ration may result in greater gluconeogenesis which can increase the possibility for lactose formation and milk production (Veen etal., 1988). This would suggest that the requirement of UIP in a ration may not be only to supply a few most limiting amino acids, but also to provide the glucogenic amino acids required for glucose production. If N available to microorganisms limits microbial growth, then there may be an advantage to increasing DIP in the rumen. However, once the microbial requirements have been met, there is no additive effect of additional rumen N (Madsen and Hvelplund, 1988). The excess DIP is converted to ammonia in the rumen and diffuses into the blood. In the same way, if the UIP requirement has been met, and no amino acids are limiting, there is no advantage to further increasing the amount of UIP in the diet beyond the effect of providing additional energy. Excess amino acids are transported to the liver, amine groups are transaminated to glutamate or arginine and eventually urea is synthesized. Understanding these possible effects and interactions of rumen protein availability makes it easier to understand why often studies fail to show positive effects of balancing rations for UIP and DIP. The requirements for each type of protein may depend on other feeds in the diet, and effects will not be seen unless the level of UIP or DIP is in fact limiting production. Some authors (Robinson and Kennelly, 1988) suggest that the requirements for UIP set by the National Research Council (1989) are Res diff UIP avai diet, UPC! more that r mquir When Why fr PTOteir. 10 too high, and in fact they are 15% higher than recommendations from the Agricultural Research Council (1984) which are used in Europe. If two diets are compared that differ slightly in the ruminally unavailable protein, and both diets supply an excess of UIP, differences in production are unlikely. Because requirements for both ruminally available and undegraded vary depending on production level and other feeds in the diet, and because methods to measure protein degradation are crude, the exact requirements of UIP and DIP are not known, and the exact amount of these protein types provided in different feeds is not known. Balancing rations for UIP and DIP is more important for higher levels of production because it is under these circumstances that ruminal synthesis lags further behind in its ability to provide for the entire protein requirement. The only way to be certain to have met the requirements for both UIP and DIP when balancing a ration is to supply a large excess of both. There are several reasons why feeding excess protein to high-producing ruminants is undesirable. Sources of protein are one of the most expensive ingredients in production-animal diets. Therefore, attempts to reduce amount of protein fed while maintaining production are important to the profitability of farms. Feeding high levels of protein results in high levels of blood urea and ammonia. These forms of nitrogen are toxic to tissues and may result in impaired reproductive performance (Visek, 1984; Ferguson and Chalupa, 1989). High protein diets were also associated with lameness in cattle (Manson and Leaver, 1988). The synthesis and excretion of urea requires expenditure of energy (7.2 kcal MB per g N) (Tyrrell et al., 1970). Finally, excess N in wastes from animal production are harmful to the natural water supply (Johnson et al., 1992). Simply feeding excess protein to ruminants is not adequate so a better understanding of the amounts of different types of protein required in rations is needed. forag. dried prote: whjcl Argo datior 1983) the ru of Up (Gian PFOIei Rendr exam I imPOr linked degfac 11 Effect of Protein Source on Degradation Feeds differ in the ruminal availability of their protein. legumes and most forages have high levels of degradable protein compared to rendering by-products and dried corn grain These feeds differ in the protein composition and interactions of these proteins with other feed components. Corn grain endosperrn proteins include glutelin which is cross-linked by disulfide bonds into a network (Duvick, 1961; Paulis, 1981; Argos et al., 1982) that explains the slow protein degradation (Harvey and Oaks, 1974). The presence of disulfide bonds has been shown to be associated with reduced degra— dation of other protein sources as well (Mahadevan et al., 1980; Wallace and Kopecny, 1983). Corn grain storage proteins (zein) are hydrophobic, and also slowly degraded in the rumen (Harvey and Oaks, 1974). While these proteins may be an important source of UIP, the lysine content (the first limiting amino acid for many rations) is low (Gianazza et al., 1977; Esen, 1980). The conformation and amino acid residues of these proteins would prevent them from coming into contact with proteolytic enzymes. Rendering by-products are composed of several types of cross-linked proteins. For example, blood meal contains the protein fibrin which is a cross-linked glycoprotein important to blood clotting, and meat and bone meal contains collagen, which is cross- linked by covalent bonding between lysine residues (Lehninger, 1975). Readily- degraded proteins typically are not cross-linked (Mahadevan et al., 1980). Proteins of plant and animal sources are frequently heated to reduce their susceptibility to degradation (Coenen and Trenkle, 1989; Aguilera etal., 1992; Nia and Ingalls, 1992). Proteins may also be cross-linked by Amadori rearrangement reactions that link lysine residues with carbohydrates (Wallace and Falconer, 1992). Heating also crosslinks proteins by formation of isopeptide bonds between the amino residue of lysine with the carboxyl residue of glutamate and aspartate (Papadopoulos, 1989). Finally, heating can result in the formation new amino acids that can be cross-linked. Fissun the am results and wi and cy: Ironica soy-b3; may re dcgrad. about t POSitivt the Pre While 5 MOWar degrad. n10m 1992). involve addif10 Soaps ( PTOVi¢ Protein degrad Crude I 12 Fissure of cystine disulfide bond may give rise to dehydroalanine which can react with the amino groups of lysine (Papadopoulos, 1989; Sathe etal., 1989). Therefore, heating results in reduced protein degradation by increased cross-linkage of proteins internally and with other proteins and with carbohydrates at specific amino acid residues. Lysine and cysteine are frequently destroyed or made indigestible from these transformations. Ironically, lysine and methionine are considered the first-limiting amino acids on com- soy-based rations fed to dairy cattle (Schwab etal., 1992), and destruction of cystein may require conversion of methionine to cysteine. If one goal of decreasing protein degradation is to increase the supply of these amino acids, heat treatment may not bring about the realization of that goal. This may be another reason some studies do not find positive results from supplementing rations with additional UIP. Mild heat treatment in the presence of sugars has been applied to reduce the destruction of amino acid protein while still protecting amino acids from microbial degradation (Mosimanyana and Mowat, 1992; Wallace and Falconer, 1992). Seed proteins are frequently treated with formaldehyde to reduce ruminal degradation of both protein and starch (Hunter and Siebert, 1987; Van Ramshorst and Thomas, 1988; Fluharty and Loerch, 1989; Hamilton et al., 1992; Oke and Loerch, 1992). The formaldehyde treatment also crosslinks protein with starch and may also involve lysine, but a better understanding of the chemical reactions is needed. In addition, proteins have also been protected from degradation by coating in calcium soaps of long-chain fatty acids (Sklan, 1989). This treatment has the dual benefit of providing increased energy to the small intestine as well as increased undegraded protein. Though most research on manipulation of the susceptibility of feeds to degradation has been with concentrates, forages provide a significant amount of the crude protein provided to high-producing ruminants. Though most forages appear to be l3 highly degraded, they differ in their susceptibility to degradation. Frequently-used tables for balancing dairy cattle rations (National Research Council, 1989) indicate that proteins of dried grasses are not as readily degraded in the rumen as those of corn silage, and that corn silage protein is less degraded than alfalfa protein. These values should be reevaluated with consideration to the techniques used to measure protein degradation of forages. The disappearance of protein from nylon bags suspended in the rumen was similar for alfalfa and corn silages, but was lower for orchard grass hay than for alfalfa hay (Janicki and Stallings, 1989). In a different study, sainfoin crude protein was degraded to a lessor extent in vitro in continuous culture than alfalfa, cicer milkvetch or birdsfoot trefoil (Dahlberg et al., 1988), but other forages did not differ from each other. Despite these effects in vitro, there were no differences when these forages were examined in situ. Silages of grasses were compared to clover silages incubated in situ (Vik-Mo, 1989). Initial disappearance of crude protein from the bags was greater for grasses, but clover silage degraded faster at later points in the incubation. Switchgrass (a warm season grass) crude protein disappeared from nylon bags more slowly than that of smooth bromegrass (a cool season grass) (Mullahey et al., 1992). A different study showed that protein disappeared faster from nylon bags in situ from the cool season grass species bromegrass and wheatgrass than for the warm- season grasses big bluestem and switchgrass (Kephart and Boe, 1991). Absolute values for protein degradation and comparison of crude protein degradation across studies is not possible due to the large variation in the techniques used to measure and to calculate protein degradation. Conservation of forages results in changes in the protein available for degradation. Frozen grass and hay crude protein (CP) was estimated to have similar amounts of rumen available CP from in situ data. but silages of the same material were estimated to have higher amounts of CP available to the rumen (Petit and Tremblay, 1992; Stalli than! prote: conse affect aCCOI McDr PrOtei PTOCe micro low p PFOdu McD( 0f the PTOteh Co'lser addilic forage: 1982). with T84 bags w. G 14 1992). The same was true for comparison of alfalfa hay and silage (Janicki and Stallings, 1989). Dried perennial ryegrass provided more CP to the duodenum of sheep than the same forage that was fed fresh or had been frozen (Beever et al., 1974). The susceptibility of protein to degradation and digestion, and the amount of protein degraded in conserved forges may be affected by various factors that affect the conservation process. Field leaf loss, rate of wilting, anaerobiosis, and heating may all affect the level of protein degradation of the conserved forages. The wilting process is accompanied by proteolysis by plant enzymes (Ohshima and McDonald, 197 8; McDonald, 1982; Petit and Tremblay, 1992). Proteins are degraded to amino acids and ammonia. Ensiling of forages results in greater solubilization and degradation of plant proteins by microorganisms (Ohshima and McDonald, 1978; McDonald, 1982). This process continues until the silage pH is reduced low enough to inactivate plant and microbial enzymes (McKersie, 1985). The only type of enzymes capable of activity at low pH, aspartic proteases, are not found in silage (McKersie, 1981). Heat can be produwd during ensiling or storage of moist hay due to respiration (Ruxton and McDonald, 1974). While heat production generally has a negative effect on the quality of the forage, some heat production may decrease the susceptibility of the forage to protein degradation (Broderick et al., 1993). Generally heat production during conservation makes the forage protein and carbohydrate indigestible and may reduce the palatability of the forage (Ruxton and McDonald, 1974; McDonald, 1982). In addition, clostridial bacteria actively degrade silage and hay proteins when these forages are not adequately preserved (Ohshima and McDonald, 1978; McDonald, 1982). Some studies have suggested that advanced plant maturity may be associated with reduced ruminal degradation of forage protein. Disappearance of CP from nylon bags was slower for silages made from more mature grasses (timothy, fescue, br pl: prt brt {Bit 863: (Ah mat pho Tron slou in p“ C0n( ferti on p Solu 1985 Van: and Wu 15 bromegrass) and legumes (red clover, birdsfoot trefoil) (V ik-Mo, 1989). Initial CP solubility was lower at advanced maturity of alfalfa, red clover, orchard grass, perennial rye grass, quack grass and timothy, but not different for birdsfoot trefoil and bromegrass (Hoffman et al., 1992). The rate of disappearance of CP from nylon bags incubated in situ of the insoluble residue was slower with advancing maturity for most plant species (Hoffman et al., 1992). Protein disappearance from nylon bags was lower in mature alfalfa samples compared to early samples (Griffin et al., 1991). Crude protein disappeared more quickly from nylon bags incubated in situ for early bromegrass compared to late bromegrass, but bromegrass maturity had no effect on N retention in non-lactating Holstein cows (Messman et al., 1992A). The in vitro degradation of CP by protease from S. griseus was affected by season for pasture samples, with the least degradation occurring in mid-summer (Abdalla et al., 1988). Several factors could contribute to this variation including plant maturity, level of regrowth (least when growth is slowest), climate (rain, temperature, photoperiod), fertilization, or plant species and variety composition. Crude protein from summer regrowth disappeared from nylon bags suspended in the rumen more slowly than spring growth of alfalfa (Griffin et al., 1991). The concentration of nitrate in plants increases during regrowth and periods where soluble carbohydrate concentrations are lower (Ourry et al., 1989). Though it would seem that nitrogen fertilization of forages would result in increased uptake of NPN and subsequent effects on protein degradation, three independent studies showed no differences in CF solubility or degradation in situ due to level of fertilization (Lindberg, 1988; Coto et al., 1989; Messman et al., 1992A). Differences in protein degradation among alfalfa varieties were demonstrated using an in vitro method with rumen enzymes (Broderick and Buxton, 1991), but were not demonstrated in a different study using the in situ technique (Kephart and Boe, 1991 ). A In treatment These pro However, protein do so a grate carbohydr; from the n degradatio greater 0p; Pla {032% P2 lignified sc Season gas for in situ d 1989). Gm bundles, am green Panic I\‘3ducti0n of Panic becaus digestible m. by Wise 1ng FOI’ag plants. Fresh nuclfiou'deS,a (Lflfleton‘ 19 l6 Intentional treatments to reduce forage protein degradation include heat treatment (Broderick et al., 1993) and treatment with formaldehyde (V ik-Mo, 1989). These procedures operate using the same principles as discussed for high-protein feeds. However, there are potentially some disadvantages of treating forages for reduced protein degradation. The forage protein is associated with high levels of carbohydrate so a greater bulk must be treated, and there is a greater potential for bonding with carbohydrate thus reducing its digestibility. Secondly, forages are believed to pass from the rumen at a slower rate. Therefore, treatments must be decrease protein degradation rate to a greater extent than is required for concentrates because there is greater opportunity for ruminal degradation. Plant anatomy may influence ruminal digestion of carbohydrate and protein in forages. Parenchymal tissues are more rapidly degraded by rumen microbes than more lignifred sclerenchymal tissues (Akin, 1979; Grabber and Jung, 1991). The warm— season grass green panic was compared in situ to the cool-season grass Italian ryegrass for in situ degradation, as studied by light and electron microscopy (Wilson et al., 1989). Green panic leaves had a larger cross section, larger and more frequent vascular bundles, and less but more densely packed mesophyll. Due to the greater rigidity, green panic leaves were broken down more by chewing than ryegrass leaves. However, reduction of the chewed leaf particles by microbes was faster in ryegrass than in green panic because ryegrass epidermis and bundlesheath cells were easily split from the digestible mesophyll cells, but green panic mesophyll was protected from degradation by these lignified structures. Forage nitrogen is found in several different forms within different structures of plants. Fresh forages contain NPN including nitrate. ammonia, amino acids, nucleotides, and chlorophyll which comprise 10 to 25% of the total crude protein (Lyttleton, 197 3). The most abundant true protein is ribulose-bisphosphate A l7 carboxylase/oxygenase (RUBP do) which comprises 30% of the total soluble true protein of temperate forages and 10 to 25% of tropical forages (Lyttleton, 1973). When isolated RUBP do was readily degraded in vitro (Nugent and Mangan, 1981). This protein was readily degraded in silage of alfalfa, orchard grass, ryegrass and fescue, but not in vetch or pearl millet (Messman et al., 1992B). The latter forages contain tannins which have been shown to bind to RUBP c/o (Jones and Mangan, 1977), and thus could in turn reduce its susceptibility to degradation . Changing the diet from fresh to dried forages was shown to decrease in vitro proteolytic activity on RUBP do of rumen fluid. This suggests that bacterial populations may adjust to the presence of the protein by changing amount or type of enzymes produced, or by changing the species of organisms, . (Hazlewood etal., 1983). Other soluble proteins are also generally enzymatic and have diverse functions (Mangan, 1982). Chloroplast membranes contain insoluble true proteins that are amphiphilic (Lyttleton, 1973; Askerlund et al., 1989; Marder and Barber, 1989), These proteins include those involved with electron transport for photosynthesis (Askerlund et al., 1989; Marder and Barber, 1989; ). Additional proteins are associated with plant cell wall (Lamport , 1980; Lee et al., 1990) or are bound to tannins (Albrecht and Muck, 1991) or lignin (Sanderson and Wedin, 1989A and B). High molecular mass phenolic compounds (tannins) bind to protein and carbohydrates to alter the ruminal availability and whole tract digestibility. Tannins may form cross-links with proteins that are hydrolyzable or condensed (Wong, 1973). Both forms of tannins would theoretically inhibit protein degradation by preventing attachment of enzymes, but cross-linkage due to hydrolyzable tannins are reversed by the acid conditions of the abomasum. Thus these tannins decrease protein degradation in the rumen but do not affect the overall protein digestion (Jones and Mangan, 1977). Other authors have correlated the concentration of tannins in sainfoin (Jones and A 18 Mangan, 1977) lotus (Barry and Manley, 1984) and across several species (birdsfoot, sainfoin, lotus, sericea) (Albrecht and Muck, 1991) with reduced protein degradation in the rumen and in silage. Tannins have also been shown to be associated with reduced protein digestion (Hagerman et al., 1992; Hanley et al., 1992). Though much of the affect of tannins on digestion is likely to be caused by the cross-linkage, tannins may also be toxic to bacteria (Hartley and Akin, 1989). Protein complexes with carbohydrates may also inhibit protein degradation. Malliard product is formed when forages are overheated during ensiling or hay-making, and this complex makes protein and carbohydrates indigestible (Goering etal., 1972; Van Soest, 1982). Glycosylated proteins are also found in fresh forages, and these linkages give the proteins unique solubility (including some TCA soluble proteins) and slower degradation (Lamport, 1980; Lee et al., 1990). Researchers have reported the accumulation of glycoproteins in the rumens of steers fed bermuda grass and alfalfa hay (Windham, 1989), which suggests they are slowly degraded. Animal and Microbial Effects on Protein Degradation The percentage protein degraded in the rumen is not completely dependent on the structural characteristics of the feed. Changes in pH could affect the solubility and conformation of proteins that may alter availability to peptide bonds for enzymatic attack (Assoumani et al., 1992). Changes in rumen reducing potential may also affect rate of protein degradation. One of the most important structural limitations to protein degradation are disulfide bonds. These bonds are disrupted and formed by non- enzymatic processes under conditions of high reducing potential. The fraction of bonds disrupted and formed at a given pH and reducing potential can be determined by application of the Nerst equation (Segel, 1976). The pH and reducing potential of the rumen varies to such an extent that as little as 10% and as much as 90% of the available disulfide bonds are likely to be disrupted on different diets (calculated from Barry et al. (l thi cle pro prot betv the t at th pepti (Barr and t Plant high sult‘h; and a 0min Prime Optim {Ound 20f“ ll 19 (1977) and Segel, (1976)). Some disulfide bonds are not likely to be accessible to the aqueous environment and will remain intact (Mahadevan et al., 1980). None—the-less, this aspect of the ruminal environment is likely to influence protein degradation. Enzymes responsible for protein degradation differ in their mode of action and cleavage sites on proteins (Barrett, 1985; Wagner, 1985). Proteases may cleave proteins at specific internal sites (endopeptidases) or at the first, second or third amino acids from the amino or carboxyl-terminus of the peptide (exopeptidases). Some proteases have both endo— and exo—peptidase activity. Proteases also cleave proteins between specific amino acids in the substrate. For example, trypsin cleaves peptides at the carboxyl group of lysine or arginine, and chymotrypsin and pepsin cleave peptides at the carboxyl group of phenylalanine, tryptophan or tyrosine. Thermolysin cleaves peptides at the amine group of 1eucine,isoleucine or valine (Lehninger, 1975). Proteases are also classified by the active sites on the enzyme into four major categories (Barrett, 1985). Serine proteases are commonly found in bacteria, plants and animals and include trypsin and chymotrypsin. Cysteine proteases are also common in bacteria, plants, and animals and include papain derived from papaya. These enzymes require high reducing potential and are inhibited by conditions that result in reaction of the sulfliydryl group of the active cysteine. Aspartic proteases are found primarily in fungi and animals and include pepsin. These enzymes have a characteristically low pH optima (pH 2 for pepsin). Metalloproteases require calcium and zinc and are found primarily in bacteria. The multitude of different enzymes that exist have different optima for pH, reducing potential, presence of inhibitors, and type of substrate (Wagner, 1985). A few studies have been conducted to attempt to characterize the proteases found in rumen fluid. Most proteolytic activity is precipitated by centrifugation at 20,000 g (Brock et al., 1982; Kopecny and Wallace, 1982). However, gentle agitation bat AC cla adt €112 vitr aCC link 6an prof or a, 20 in phosphate buffer resulted in solubilization of much of this activity from rumen bacteria (191% of original) without disruption of cells (Kopecny and Wallace, 1982). Additional activity of similar character was removed by blending and treatment with detergent (Kopecny and Wallace, 1982). Cysteine proteases were the most dominant class found in rumen fluid (Brock et al., 1982; Kopecny and Wallace, 1982) , with additional activity of serine and metallo-proteases (Brock et al., 1982). Rumen enzymes included a broad range of activities on different synthetic substrates including both endo— and exoprotease activity (Wallace and Kopecny, 1983). None—the-less, in vitro (Wallace, 1992) and in vivo (Chen et al., 1987) studies demonstrated an accumulation of hydrophobic peptides from incomplete protein degradation, which demonstrates that protease activity in the rumen is limited in the ability to cleave some linkages. Protein degradation in the rumen may depend on changes in the rumen environment that affect the different proteases, and on the changes in the protease profile found in the rumen. The amount of the different types of proteolytic enzymes, or the amount of substrate, can affect the fraction of available substrate that can be degraded according to Michaelis—Menten kinetics (Lehninger, 1975). Different protein sources resulted in different determinations of Km and Vmax for the same rumen fluid samples (Broderick and Clayton, 1992). Therefore, some feeds not only may be more likely to degrade at a faster rate than other feeds, but the susceptibility of some feeds to enzyme limitation of the rumen fluid may be greater than for other feeds. In other words, some proteins may degrade more uniformly when fed to different animals and across diets than other proteins. The level of protein degradation activity on RUBP do was more than doubled by changing the ration from hay to fresh forage (Hazlewood et al., 1983). This would indicate a possibility for adaptation of rumen organisms to the protein source. A 11.7%? fig—"A13 ":7 . ' W anin iono‘ nesul data t degra increa Proleir Vary dc There fc Protein 21 different study showed that feeding the resistant protein albumin resulted in increased numbers of proteolytic bacteria (Butyrivibiro spp.) and increased activity on azocascin, but the activity on albumen was not increased relative to that of azocascin (Wallace et al., 1987). Therefore, fluctuations in proteolytic activity may exist, but evidence of adaptation to specific proteins was not demonstrated. A lower level of ruminally degraded carbohydrate in the ration was shown to decrease protein degradation in vivo (Meyer et al., 1986; Cecava et al., 1988) and in situ (Meyer et al., 1986). However, protein disappeared faster from nylon bags incubated in rumens of sheep fed grass than those fed barley (Ganev et al., 1979). Proteolytic activity measured in vitro was higher when rumen fluid was collected from animals fed high energy diets (Furchtenicht and Broderick, 1987). The addition of the ionophore monensin, which inhibits the proteolytic organism Streptococcus bovis, resulted in reduced ammonia production in the rumen (Chen and Russell, 1989). These data would suggest that proteolytic enzyme activity in the rumen partially limits protein degradation, because diets that increase this activity, such as those high in starch, increase the portion of feed protein degraded. The percentage of protein degraded in the rumen depends upon the rate of protein degradation and on the rate of passage of the protein from the rumen. The faster the protein passes, the less time is available for protein degradation. Passage rates will vary depending on rumen size, diet, intake level, and nature of the feed protein. Therefore, undue rigidity should not be placed on values of expected UIP for a given protein source. Methods to Determine Protein Degradation Measurement of Protein. Ruminant nutritionists have developed the practice of measuring nitrogen instead of true protein. Forage protein is about 16% nitrogen, so the estimation of protein in forages is 6.25 (x) nitrogen content (Association of Official (H mt re; Fre 8CC hav deb: 197t bind eitht With men 198: men Plot. mea in Vi mlCr 22 Agricultural Chemists, 1975). The macro-Kjeldahl method is commonly used to measure feed nitrogen (Association of Official Agricultural Chemists, 1975). Recent modifications have made the method faster and reduced the requirements for reagents (Hach et al., 1985; Schmitter and Ribs, 1989). Comparison of traditional Kjeldahl method with use of block digestion or peroxide catalysts showed that there was good repeatability and recovery of nitrogen when using the peroxide method, but some samples were not adequately digested in block digesters (Watkins et al., 1987). Frequently protein is measured by spectrophotometry after reaction with folin-phenol reagent (Lowry et al., 1951). However, several compounds in forages interfere with accurate measurement of true protein by this method (Askerlund et al., 1989). Attempts have been made to separate interferences from protein by TCA precipitation before determination of protein degradation by Lowry's method (Bensadoun and Weinstein, 1976). Attempts to measure forage protein by spectrophotometric determination after binding with bicinchoninic acid (BCA) were not well correlated with Kjeldahl nitrogen either before or after precipitation with TCA (Messman and Weiss, 1992). Interference with forage phenolics or other compounds was suspected as the problem. Estimation of Microbial Protein Synthesis. Several reviews have described methods used to measure protein degradation of feeds (Orskov et al., 1980; Lindberg, 1985; Madsen, 1985; Nocek, 1988; Michalet-Doreau and Ouldbah, 1992). These methods can be confined to one of three categories: 1) in vivo measurement of feed protein arriving at the small intestine by collection from duodenal cannulas, 2) in situ measurement of protein disappearance from nylon bags suspended in the rumen, and 3) in vitro laboratory methods. One common difficulty for methods that involve live microorganisms is the estimation or separation of microbial protein from undigested feed protein. A E) S)". 19 ma ma prc (H: “281 alsc 23 Various procedures have been used to determine the proportion of microbial N associated with feed in digesta. These procedures include measurement of modified amino acids or sugars that are found exclusively in microorganisms such as diaminopimelic acid (DAPA) (Czerkawski and Foulds, 1974), arninoethylphosphoric acid (AEPA) (Ling and Buttery, 1978), D-alanine (Rooke et al., 1984) or L-glucose (Rooke et al., 1984). Total RNA or purine concentration also is used with the assumption that RNA of feed origin is entirely digested (Zinn and Owens, 1986). Excreted metabolites of nucleic acids have also been used to estimate microbial synthesis in less intrusive studies ( Lindberg et al., 1989; Lindberg and Jacobsson, 1990; Puchala and Kulasek, 1992). The most problematic aspect of using these internal markers for estimation of microbial growth is the lack of uniformity of concentration of marker per amount microbial nitrogen across the microbial ecosystem. For example, protozoa have no DAPA and have higher concentrations of AEPA titan bacteria (Henderickx et al., 1972; Ling and Buttery, 1978). Whereas, DAPA is associated with the cell wall of gram positive bacteria, this concentration will vary with conditions that result in changes in the type of microbial growth. The RNA content of bacteria may also vary depending on the microbial population and rate of growth (Matin, 1992). Stable and radioactive isotopes of nitrogen (Nolan et al., 1972; Varvikko and Lindberg, 1985; Javorsky et al., 1986; Aharoni et al., 1991), sulfur (S) (Henderickx et al., 1972; Raab et al., 1983; Spencer et al., 1988; Thomsen, 1985), or phosphorus (P) (Buckholz and Bergen, 1973; Smith etal., 1978) have been used as microbial markers but radioisotopes may require costly safety procedures for large animal studies and mass spectroscopy for 15N may be expensive or inaccessible. Inorganic forms of the nutrient such as nitrate (N03), phosphate (P04 ) or sulfate (S04) often are placed in the rumen or in vitro flask and microbial growth is estimated based upon isotope A 24 incorporation into microbial matter. Isotopic methods label both bacteria and protozoa without labeling feed sample. Several studies have compared the use of different microbial markers on the estimation of microbial CP. Some authors have reported similar estimates of microbial CP from estimates made using DAPA vs. RNA (McAllan and Smith, 1983), or DAPA vs. 358 (Mercer et al., 1980). Since DAPA is to estimate only bacterial protein one would expect it to underestimate the total microbial CP. However, DAPA predicted much greater microbial CP than 15N (Varvikko and Lindberg, 1985), or higher and unrealistic estimates compared to 35S (W hitelaw et al., 1984), and it resulted in an estimate that was nearly twice as high as from 15N or 358 (Siddons et al., 1979). Other work did show DAPA to underestimate microbial CP compared to 35S and RNA with an important diet by marker interaction effect (Ling and Buttery, 197 8; Smith et al., 1978). Microbial CP estimated using 32F to pass to the duodenum was 85% the level estimated from RNA (Smith et al., 1978). Because 32? labels the nucleic acids synthesized by bacteria, the authors concluded that differences in microbial yield determined with the two markers resulted because some feed RNA (which would not be 32P labeled) was not digested in the rumen. The assumptions and problems of using different external markers are similar, but further discussion will focus on use of 35S because this marker was employed in a study described in this dissertation. Radioactive sulfur often is used owing to its low energy B-emission which reduces the risk of hazardous radiation exposure. Various microorganisms bring about the enzymatic reduction of S04 to sulfide (H2S) which is then converted to the thiol group of cysteine by reaction with pyruvate and ammonia or alternatively with serine. Methionine is synthesized from the cysteine by rumen microbes. Therefore, radioactivity associated with protein after incubation of microbes with inorganic sulfur results from microbial synthesis of amino acids. 25 A ratio of nitrogen to radioactive sulfur (Nz35S) can be determined for a standard which is entirely of microbial matter. This ratio can then be multiplied by the amount of 358 associated with a mixture of dietary and microbial protein. The product is the estimate of microbial N. The standard of 100% microbial protein may be obtained by isolating protein from an in vitro rumen incubation that used only non— protein nitrogen (Henderickx et al., 197 2). Alternatively, microbes may be separated from dietary protein by differential centrifugation (Walker and Nader, 1975). Another alternative to evaluating the N2353 ratio is to hydrolyze total protein and isolate total methionine or cysteine, then quantify percent microbial protein as 35S-methionine or cysteine/total methionine or cysteine x 100 (Gunn and McMeniman, 1989). This approach is obviously time consuming, expensive and still requires assumptions about the sulfur amino acid content of microbial and partially digested protein. The successful use of 358 as a microbial marker requires the validity of a several assumptions. Inorganic sulfur must be the only source of sulfur for methionine or cysteine incorporated into microbial protein. If sulfur amino acids from the feed protein source itself are used instead of inorganic sulfur, total microbial protein will be underestimated. In vitro incubation of rumen fluid containing 358 and 14C—labeled sulfur amino acids as a nitrogen and energy source resulted in less than 1% and 11% incorporation of free cysteine and methionine respectively directly into microbial protein (Nader and Walker, 1970). The majority of the microbial sulfur was provided from the inorganic sulfur pool which resulted in low 14C:35S ratio of microbial sulfur amino acids. The authors concluded that direct incorporation of sulfur amino acids from feed sources was negligible. Results obtained from a similar in vitro procedure indicate that the form and amount of 358 used is an important consideration (Halverson et al., 1968). Cystine and methionine added to a sulfate and rumen fluid incubation appeared to be more readily 26 converted to H2S than 358-1abeled 804. Moreover, added H28, cystine and methionine depressed the incorporation of 35S-labeled 804 into microbial protein though casein had no consistent effect. The authors attributed the decrease in 353- labeled 804 incorporation into microbial protein to incorporation of H28 from cystine and methionine into the protein in place of H28 from 804. These results suggest that 804 may not be an adequate carrier of 358, and that labeled H28 or labeled cysteine are preferred. An additional assumption that must be true for accurate measurement of microbial contamination is that microbial proteins produced from different types of fermentation have the same N:358 ratio. It is particularly important that the isolated microbial cells or microbial protein used as a standard have the same N:358 ratio as the sample of unknown quantity of microbial protein. The N:358 ratio of microbial culture in NPN media may differ from that of the culture in media with true protein to assay. Non-protein components of the feed sample might further alter the microbial metabolism causing a difference in N2358 ratio between standard and samples and among different samples. Addition of excess energy and microbial growth factors (vitamins, minerals, isoacids) may diminish effects of non-nitrogen feed components on microbial action. The standard used to determine Nz358 ratio alternatively may be obtained from differential centrifugation rather than culturing microbes in non-protein nitrogen media. Methods of differential centrifugation may isolate only certain microbes that have different Nz358 ratios than the overall microbial population. For example, bacteria bound to fiber and protozoa may not be recovered by such separation and may differ in sulfur-amino acid content or in their ability to incorporate inorganic sulfur. In addition, the cell fraction recovered may not have the same S-amino acid content as whole cells. Bacteria have virtually no S-arnino acids in cell wall (Mathers and Miller, 1980). and A _.._. —» 7' f»? hrrigeflf- 1‘ , .7 27 nucleic acids represent 30% of the total N content of microbes though they contain no sulfur. Bacteria are 52% true protein and 20% RNA and DNA (Gottschalk, 1985). RNA is 18% N and protein is 16% N (calculated from Lehninger, 1975). The behavior of inorganic sulfur during fermentation adds complexity to the use of sulfur as a microbial marker. Elemental sulfur (SO)is insoluble in aqueous solutions. At pH 7 and Eh -.25 V (possible conditions in the rumen) the ratio of [80]:[H28] is calculated from the Nerst equation to be .46. Greater reducing potential and lower pH reduce this ratio exponentially so that if pH is reduced to 5 or Eh is reduced to -.35 V elemental sulfur is less than 1% of H28 (Segel, 1976). Under reducing conditions and low pH, H28 gas may be evolved after exposure to air (Walker and Nader, 1968), and it may be oxidized rapidly to 80. Oxidation of samples to convert 80 solid to 8042' has been used to adjust samples before analysis. None-the- less, the volatilization of inorganic sulfur compounds results in additional difficulty in use of sulfur as a marker, especially for in situ or in vivo studies (Mathers and Miller, 1980). These possibilities necessitate careful technique to maintain a closed environment, and an excess of inorganic sulfur may be necessary, as well as appropriate interpretation of results. Another difficulty related to solubility of sulfur compounds is that as much as 20% of inorganic sulfur was found to associate with protein added to sterile rumen fluid (Walker and Nader, 1968). In this experiment sulfur was precipitated with barium chloride (BaCl2) after digestion in nitric acid (HN03) and perchloric acid (HClO4.). Though the samples were acidified before digestion, some elemental sulfur (80) may have been associated with the protein especially if oxidation occurred before acidification . Then 80 would have been converted to sulfate (8042‘) during the oxidative digestion. In order to eliminate inorganic sulfur associations with protein other workers hydrolyzed protein in HCl after oxidation with perforrnic acid (H2CO3) _ —____. . wizfi 28 and prior to precipitation of sulfate using BaCl2 (Mathers and Miller, 1980). The perforrnic acid oxidation converted S0 to 8042'. In Vivo Methods to Measure Protein Degradation. Determination of protein degradation in the rumen may be accomplished by collection of digesta samples from surgically placed cannulas in the abomasum (McAllan and Smith, 1983) or duodenum (Santos et al., 1984; Van Bruchem et al., 1988; Gasa et al., 1991). Duodenal cannulas are placed in the proximal region to minimize intestinal absorption before sampling and to prevent secretions from the pancreatic and bile ducts from contaminating the sample. Reentrant cannulas are occasionally used, and they can be automated so that total digesta is diverted outside the intestine from which it is sampled and returned. More commonly, and with less physical stress and complications, simple "T" cannulas are used. From these cannulas, samples can be taken during desired periods of time. A stop-gate mechanism can be inserted to prevent a separation of digesta that results in the sample being of a different composition than the digesta which passes beyond the cannula during sampling. None-the—less, it is conceivable that fluid digesta will be collected in greater quantity than particulate matter, because liquids are more likely to flow from the cannula after it is opened than are solids. Samples of digesta from simple cannulas therefore may not be representative of the digesta passing from the rumen. Use of duodenal cannulas to measure protein degradation requires use of indigestible markers to determine total digesta flow in order to determine the proportional sample size. The use of these markers has been reviewed (Faichney, 1975). An indigestible marker must be administered continuously until steady state conditions are achieved. The passage of digesta (L/d) can be determined as flux of infusion of marker (g/d) divided by the marker concentration in digesta (g/L). Marker infusion may be discontinued after steady state conditions are achieved. The passage is then determined from the decline in marker concentration from the first order equation 29 -d[A]/dt = k[A], or more directly from the integral of this equation [A]i = [A]0 (x) e'kt, whereas d[A]/dt is the fractional change in marker concentration, k is the fractional rate of passage, and [A] is the concentration of marker initially ([A]0) or at time t ([A]i). When using simple cannulas that result in unrepresentative sampling of duodenal contents, a two-marker method is required to correct for the disportionate amount of fluid phase relative to particulate phase of digesta collected (Faichney, 1980; Faichney, 1992). The currently available markers have various limitations that result in inaccuracies of the passage estimates. Markers have been shown to be partially digested and absorbed, to dissociate from the phase they are intended to mark, or to bind to the gut wall or to other feeds (Warner and Stacy, 1968; Faichney, 1975; Teeter and Owens, 1983; Cruickshank et al., 1989). These limitations as well as the difficulty of using markers correctly add to the inaccuracy and impreciseness of in vivo methods. Estimation of protein degradation in vivo is made by measuring the concentration of protein in samples collected from the small intestine. This value (g 1 protein lg DM collected) is multiplied by the total passage of digesta through the rumen , determined by markers (g DM / d). From this value for total protein flow to the small intestine, microbial protein is subtracted. Microbial protein is determined by using 8 microbial markers with all their limitations as discussed previously. The passage of undigested feed protein per day is then divided by the total protein intake per day to determine the fraction of undegraded protein in the diet. In vivo methods are certainly _ more physiological than other methods but they potentially include great biases and variation (Nocek, 1988). In Situ Methods to Measure Protein Degradation. The most common means to estimate protein degradation of feeds is by submersion of feed samples in mesh bags . in the rumen of cattle or sheep for different durations in time followed by determination of residual N. The protein source is often divided into at least 3 fractions for that CXIEI ofth previ rum 6! andth €Stimal 30 description of its disappearance (Mohamed and Smith, 1977; Nocek, 1988). The "A" fraction is that which is washed from the bags by rinsing before submersion in the rumen. This protein is considered degraded in the rumen. The "B" fraction is protein that is slowly and partially degraded in the rumen, and partially digested and absorbed in the small intestine. The "C" fraction is protein which remains in the nylon bag for extended periods of time and which is not digested. The first order rate of degradation of the B fraction is determined from in situ data using the model described for passage previously. The portion of the B fraction of dietary protein calculated to be degraded in the rumen is determined from the in situ disappearance rate divided by the sum of this rate and the estimated passage rate (Qtskov and McDonald, 1979; Kristensen et al., 1982). Often the same passage rates are assumed across diets and feed types, while other estimates are based on different passage rates for different feeds (Ganev et al., 1979). Passage rates are likely to vary across feed types and within fractions of the same feed, 1 but determination of the passage rates for the feeds on different diets requires passage markers which are subjected to the same difficulties as described for in vivo studies. , Additional calculations include allowance for a lag phase before digestion of sample , (McDonald, 1981; Dhanoa, 1988), but if lag phase is assumed to affect both digestion and passage equally it can be excluded from the calculation (Allen and Mertens, 1988). i In situ N disappearance is affected by bag porosity, particle size of sample, and sample size (Nocek, 1988). This disappearance is also affected by the diet fed to the ‘ host animal (Murphy and Kennelly, 1987; Caton et al., 1988). While these effects make it impossible to compare feeds for degradation across studies, protein degradation 1 of a feed is not entirely a function of the feed, but is also affected by the ruminal conditions, so variation across methods is expected and is interesting in itself. More troublesome aspects of the in situ method do exist. The pH inside the nylon bag has fn sitl CO! mmen Control animal. Unifonn fraCLion . and how I degMdatic 962mm,“ ref"(Isifltlptt 0"antenna 31 been shown to be lower than that outside the bags, especially when small pore sizes were used (Marinucci et al., 1992). The microbial population inside the bags also differed from that outside the bag. Both protozoa and bacterial populations were lower inside the bags (Meyer and Mackie, 1986; Marinucci et al., 1992). Analysis of digesta from nylon bags incubated in vitro showed that proteins escaped the nylon bags before being digested (Spencer et al., 1988). The microbial attachment to feed incubated in situ is frequently not measured, though several studies have shown high levels of contamination (Nocek, 1988). In Vitro Methods to Measure Protein Degradation. Laboratory methods to determine protein degradation range from in vitro fermentation with markers of microbial growth to prediction of degradation from protein solubility. In vitro techniques are less expensive than in vivo and in situ methods, and they offer the possibility to analyze and compare several feeds at once. In vitro techniques also allow the researcher to collect and analyze botlt the residue and the metabolites of microbial degradation, while more physiological techniques are complicated by fluxes out of the rumen or out of in situ incubation bags. In vitro methods may ultimately allow for control of the various non-structural factors that alter degradation of a feed (microbial, animal. environment) and therefore allow for the possibility to characterize feeds uniformly for their susceptibility to CP degradation. As with other methods, the fraction of protein degraded in the rumen requires estimation of rates of passage The in vitro environment for ruminal fermentation studies can vary substantially and how this variation occurs and its effects must be understood. The rate of in vitro degradation of protein depends on the amount of protein in an accessible form and on the amount and types of enzymes and bacteria present. Non-structural aspects of the feed sample itself (protein and non-protein components) may affect the type and extent of fermentation and therefore affect the extent of microbial degradation. For example, A 32 if a feed sample does not contain adequate ferrnentable carbohydrate or is deficient in an essential mineral, poor microbial growth may limit protein degradation regardless of protein structure. High-producing dairy cattle are fed mixed rations to meet the requirements of both the animal and its microbial population. This high plane of nutrition results in extensive fermentation and presumably abundant microbial enzymes in vivo. Therefore the extent of microbial degradation of protein in vivo is likely to be limited by the physical nature of the protein itself and by its surrounding feed components, rather than by the ability of the organisms to grow on the material. If an individual feed is only one component out of several in a feeding regime, an appropriate in vitro system would supply other nutrients besides the sample to be tested in order to insure profuse microbial growth relative to the amount of sample. The resulting protein degradation then would mainly depend on the structural nature of the feed. Fermentation in continuous culture systems allows removal of metabolites that can reduce activity of fermentation (Dahlberg et al., 1988; Fuchigarni et al., 1989). Incubation in batch culture is more convenient, especially if several treatments are to be compared. These systems require added buffers to maintain adequately high pH, but harmful metabolites are not removed. If a system contains an abundance of microbial nutrients, microbial growth may be limited by accumulation of volatile fatty acids (VFA) or low reducing potential or pH. Therefore the amount of protein used in an in vitro system for protein degradation must be low enough so that growth is limited by ability of the microbes to digest the protein, and not by a reduction in pH or accumulation of VFA's. The importance of available protein for microbial growth using the system described above implies that not only is protein degradability (or lack thereof) likely to limit microbial growth (and therefore degradation), but also certain amino acids may be limit requi prote deten Micrt (Raal differ level (Wall Walla mark: dflsire M6831 marks: high 1, Showe Protei. 1985)_ and di: with m 33 limiting. Isoacids of branched-chain amino acids can be added to meet some of these requirements so that limitations on extent of degradation will be due primarilly to protein structure and its associations with other feed components and not due to its amino acid profile. Most in vitro fermentation methods use markers of microbial growth to determine microbial contamination of feed sample. Since these markers are difficult to use and the results are unreliable, attempts have been made to circumvent their use. Microbial growth has been estimated from turbidity of fluid or from gas production (Raab et al., 1983). These approaches provide crude estimates with potential for bias as different fermentation patterns result in different levels of gas production at a given level of microbial protein synthesis. Proteins have been labeled themselves with 14C (Wallace et al., 1987; Wallace and Falconer, 1992) or azo-dye (Mahadevan et al., 1979; Wallace and Kopecny, 1983) for determination of degradation from the release of marker. This approach is useful in some situations but is more labor-intensive than desired. The degradation of individual proteins has been followed by separation of these protein from microbial proteins using electrophoresis (Fahmy et al., 1991; Messman et al., 1992B; Wallace, 1992). A possible future approach to measuring protein degradation without microbial markers, but in the presence of high levels of carbohydrate, is to incubate feeds with high levels of corn starch and cotton but with low levels of feed protein. Previous work showed that under these conditions starch was optimally degraded with urea as a protein source, but that peptides were required for degradation of cellulose (Thomsen, 1985). Therefore, disappearance of starch may be a predictor of ammonia availability and disappearance of cellulose may be related to amino acid availability of feeds. Another approach to measure protein degradation has been to incubate feeds with rumen fluid under anaerobic conditions without addition of a carbohydrate source 34 (Russell et al., 1983). Rumen microbial growth was inhibited this way and therefore was not required to be quantified (Chen and Russell, 1989). A theoretically similar technique is to incubate feeds in rumen fluid in the presence of inhibitors of N uptake by microbes (Broderick, 1987; Neutze et al., 1993). The most serious limitation of these methods is that enzyme activity is not likely to be in excess of substrate. For this reason, higher level of substrate in the incubation with rumen fluid and inhibitor resulted in reduced fractional degradation of CP (Broderick and Clayton, 1992), and level of activity depended on rumen fluid sample (Furchtenicht and Broderick, 1987). Ruminal proteolytic enzymes have been shown to be loosely associated with bacterial cell wall, and can be solubilized while retaining activity by gentle extraction (Kopecny and Wallace, 1982; Brock et al., 1982). A procedure to compare protein degradation of feeds from crude extracts of rumen enzymes has been published (Mahadevan et al., 1987). Using this method, relative differences in liberation of amino acids from common feeds were realistic, but the fractional CP degradation was much lower than ‘ expected. This would suggest that enzyme was more limiting in this system than would 1‘ be in vivo, or some other aspect of the in vitro system (i.e. lack of microbial attachment) limited degradation. 1 The difficulties associated with fermentation studies, including standardizing the fermentation, measurement of microbial growth, and the requirement of access to an ‘[ inocula source (fistulated ruminant), have led to the development of several methods to imeasure protein degradation using enzyme preparations. Ruminal protease was enriched from Bacteroides amylophilus and used for studies of degradation of several :protein sources (Mahadevan et al., 1980). Alternatively, methods have used commercially-available proteases. Crude protein degradation using five different commercially-available proteases was compared to that of the in situ method for several concentrate feeds (Poos-Floyd et al., 1985). The solubility of CP was determined by filter and l diffe data. feeck degrz 1983 both Tetra prote- 1992] lheol exist i Dossit exist i llunch en 2m 35 filtering after degradation with: Streptomyces griseus protease, papain, bromelain, ficin and Aspergillus oryzae protease. Though absolute degradation with proteases was different than observed in situ, all enzymatic degradations were correlated to in situ data, suggesting the possibility for using enzymes to detect relative differences among feeds. Use of S. griseus enzyme has become popular for prediction of protein degradation of feeds but the optimal pH f0r the enzyme is 8.0 (Krishnamoorthy et al., 1983), which suggests a type of activity not found in the rumen. It has been used at both normal rumen pH with reduced activity (Chamberlain and Thomas, 1979; Terramoccia et al., 1992) and at the higher pH (Krishnamoorthy et al., 1983). A neutral protease was recently introduced to measure protein degradation (Assoumani et al., 1992). The use of commercial enzymes has not provided consistent results. Theoretically, there may be a problem that only one of the many types of proteases that exist in the rumen is present in a purified enzyme. More serious, however, is the possibility that the type of protease activity present in the purified enzyme does not exist in the rumen. A comparison of protein degradation by crude extract from the rumen and by S. griseus protease showed very different patterns of degradation across several feeds (Mahadevan et al., 1987). In addition, proteases may not be the only enzymes required for degradation of plant proteins. Structural limitations from carbohydrates could prevent degradation by protease enzymes. Crude protein degradation was lower when incubated with 8. griseus protease compared to in situ methods, which suggests a synergism of protease with amylase and cellulase (Terramoccia et al., 1992). This difference was especially apparent for forages (J anicki and Stallings, 1989). Addition of amylase and beta-glucosidase increased protein degradation with neutral protease (Assoumani et al., 1992). 36 Various researchers have attempted to characterize feed protein based upon solubility properties in aqueous solutions such as saline, buffers, or autoclaved rumen fluid (Wohlt etal., 1973; Pichard and Van Soest, 1977; Crawford et al., 1978; Crooker et al., 1978; Mahadevan et al., 1980; Krishnamoorthy et al., 1982; Nocek et al., 1983; Stern and Satter, 1984). Nitrogen solubility varies greatly for different feedstuffs. For example, while nitrogen of brewer's grains was only 3% soluble in borate-phosphate buffer, oat nitrogen was 55.4% soluble (Krishnamoorthy et al., 1982). Buffer soluble nitrogen usually is comprised mostly of non-protein nitrogen (NPN), such as ammonia, urea, nitrates, amino acids and small peptides (Pichard and Van Soest, 1977; Krishnamoorthy et al., 1982). Nucleic acid nitrogen is the major NPN fraction not soluble in neutral buffer, but generally it is low in quantity, and is underestimated by nitrogen analysis using the Kjeldahl procedure (Lyttleton, 1973). In addition to NPN, some true protein is soluble to varying degrees among feeds. Protein solubility can be influenced by various factors associated with the solvent or extraction procedure. Subtle changes in chemical composition of solvent can have pronounced effects on nitrogen solubility. For example, substitution of ammonium chloride with sodium chloride in Burrough's mineral mixture (BMM) resulted in increased nitrogen solubility for several concentrates (Crooker et al., 197 8). It is difficult to draw meaningful conclusions, however, as to the effect of solvent composition on nitrogen solubilization because substantial interaction among feedstuffs, methodologies and solvents exists. For example, in one study, nitrogen solubility was found to be higher in BM or .15 M sodium chloride (NaCl) than in autoclaved rumen fluid (ARF) (Crawford et al., 197 8), but other authors have reported the opposite to be true (Waldo and Goering, 1979), and still others found no significant difference among the three solvents (Lamport, 1980). It is noteworthy that similar concentrate feeds were analyzed in each of the above studies. nit ion: 37 Comparisons of ARF, NaCl, BMM and hot water revealed that variance among feeds was most pronounced for ARF, suggesting that this method may best separate different feeds according to solubility of protein (Waldo and Goering, 1979). However, while homogeneous variance within samples was detected in the above experiment (Waldo and Goering, 1979), the use of ARF may not be repeatable from laboratory to laboratory, or from time to time, as ARF composition varies (Jancarik and Proksova, 1970). Solvents often are evaluated according to their similarity to nitrogen solubilized in ARF (Krishnamoorthy et al., 1982). Ironically, solubility in BMM was found to be more highly correlated to nitrogen disappearance from nylon bags than was solubility in ARF (Craward et al., 1978). As with ARF, BMM has a short shelf life and contains nitrogen unless it is modified (Krishnamoorthy et al., 1982). Purified proteins are insoluble in isoelectn'c solutions, or those in which the ionic concentration is such that the protein's positively charged molecules are equaled by its negatively charged molecules (Lehninger, 1975). Ionic concentration of NaCl was demonstrated to effect solubility of soybean meal proteins (Smith et al., 1959). Only 45% of soybean meal nitrogen was soluble at .1 ionic strength, while 98% was soluble in distilled water, and about 80% was soluble in .5 ionic strength NaCl (interpreted from graph) (Smith et al., 1959). These results were not confirmed, however, when seven other concentrates were compared for solubility in BMM, NaCl and modified BMM, where NaCl was substituted for ammonium chloride, with ionic strengths of .11, .15 and .19 (Crooker et al., 1978). The narrow number and range of ionic strengths studied may be responsible for this finding. Ionic strength of rumen fluid was calculated as .15 (Salobir et al., 1970). The pH of a solvent has been shown to influence nitrogen solubility of concentrate feed protein (Smith et al., 1959; Wohlt et al., 1973; Waldo and Goering, 1979). Eighty-five percent of soybean meal protein was soluble in aqueous HCl (pH 38 2.0) and water (pH 7.2), while less than 10% was soluble at pH 4.0 (Smith et al., 1959). Significant differences in solubility of isolated soy protein and casein also were observed when pH of autoclaved rumen fluid and BM were varied from pH 5.5 to pH 6.5 (W ohlt et al., 197 3). Therefore, the effect of pH on protein solubility is an important consideration when choosing an appropriate solvent. Ironically, solubility often is measured in neutral pH solutions, even though rumen fluid is acidic to varying degrees. Whereas sodium chloride has almost no buffering capacity, and bicarbonate- phosphate buffer has an unstable pH, a borate phosphate buffer has been suggested as an appropriate substitute to measure nitrogen solubility without variation due to fluctuation in pH (Krishnamoorthy et al., 1982). Solubility of feed nitrogen as a function of time is important in relating data to the dynamic system of the rumen. Feeds vary in the rate of solubilization, and these rates have been interpreted as single (Mehrez and Qrskov, 1977) and multi-exponential curves (Waldo and Goering, 1979). Whereas plant enzymes may be active, solubilization over extended periods of time may be due to degradation of protein instead of simple solubility. Buffer-soluble CP appears to be more readily degraded in the rumen than insoluble nitrogen. There is a close association between buffer-soluble CP and nitrogen disappearance from nylon bags suspended in the rumen, especially when feeds are incubated in situ for short periods of time, such as 1 h (Nocek et al., 1979; Stern and Satter, 1984). In addition, the amount of buffer soluble nitrogen in a feed was correlated strongly to increases in rumen ammonia concentration after feeding (Jones and Mangan, 1977; Waghom, 1991). However, other authors found no consistent relationship between nitrogen solubility in various aqueous solvents and ammonia concentration (Crooker et al., 1978). Nitrogen which is available to microbes is converted to ammonia is used for microbial synthesis of true protein. Soluble NPN is 39 rapidly utilized by rumen bacteria, and solubility in mineral buffer of pure proteins was correlated strongly to degradation rate in rumen fluid (Hamilton et al., 1992). Based on these findings, one may suspect that buffer-soluble nitrogen predicts immediate availability of nitrogen to rumen microorganisms for a given feed, provided that soluble nitrogen is primarily NPN or is easily—degradable true protein. This assumption would appear to be true for most of the feeds that have been analyzed. However, rate of degradation of soluble true protein varies among sources, and a few feeds have been found to release some soluble protein that is not rapidly degraded (Nocek et al., 1983; Mahadevan et al., 1980). For example, when soluble and insoluble feed fractions from dehydrated alfalfa and corn gluten feed were subjected to Streptomyces griseus protease, amino acid release rates were only half as great for soluble fractions as for insoluble fractions of the same feeds (Nocek et al., 1983). Therefore, one may wish to determine the amount of soluble true protein that comprises the soluble nitrogen fraction for some feeds. This task is easily accomplished by protein precipitation from solution with such reagents as trichloroacetic acid (T CA) (Bensadoun and Weinstein, 1976). The rate of degradation of soluble protein has not been determined for many feeds. This observation especially is true for forages. The major soluble protein of plant leaf material, ribulose bisphosphate carboxylase/oxygenase (RUBP-Clo) or fraction I protein, has been shown to be readily degradable in vitro and in vivo when derived from fresh forage (Nugent and Mangan, 1981; Hazlewood et al., 1983). However, it may bind to tannins or other forage components or be altered during forage conservation causing it to become less readily available in other circumstances. Some authors have reported a poor correlation between nitrogen solubility in buffer with in vivo protein degradation across several feeds (Stern and Satter, 1984). This result is not surprising given that the amount of soluble feed protein appears to 40 have no relationship to the degradation rate of the remaining insoluble protein fraction of a given feed (Mahadevan et al., 1980; Nocek et al., 1983). Therefore, one can conclude that estimation of nitrogen solubility in neutral buffers may predict amount of nitrogen immediately available to bacteria, but it may not be accurate for all feeds as discussed above, and nitrogen solubility in aqueous solutions is not sufficient to determine protein degradation rate or protein degradability. Crude protein solubility in detergents has also been studied briefly. The amount of pH-7—detergent insoluble hay protein was negatively correlated (R = -.83) to in situ degradation, but the correlation for silages was lower (R = -.51). Crude protein insoluble in pH 7 detergent was shown to be slowly degraded in the rumen (Lindberg, 1988). Heat treatment has been shown to increase the detergent-insoluble crude protein as well as protect forages from ruminal degradation (Weiss et al., 1989). This fraction also increased with advancing maturity, and was more concentrated in stems than leaves of alfalfa (Sanderson and Wedin, 1989). Detergent (pH 7) insoluble protein was also shown to be higher in pasture samples that had higher concentrations of S. griseus- protease-resistant protein (Abdalla et al., 1988). Crude protein that is insoluble in detergent in 1 N H2804 (ADCP) has been associated with indigestibility of protein and carbohydrate. Severe heat damage renders protein indigestible by non-enzymatic browning (Maillard) , and the product of this reaction is insoluble in sulfuric acid-detergent solution (Weiss et al., 1986A; Weiss et al., 1986B; Van Soest, 1965). The amount of ADCP and pepsin-insoluble CP were correlated across a large range of heat-damaged alfalfa, but with an R2 of only .66 (Goering, 1972), though it was well correlated with apparent in vivo indigestibility of forages (Goering et al., 1972). Some ADCP is digestible so it is not a uniform fraction in forages or concentrates (Britton et al., 1986; Weiss et al., 1986; Waters et al., 1992; Broderick et al., 1993). Theoretically, Maillard product from short peptides and 41 saccharides could also be filterable yet indigestible, and so added error may be associated with use of ADCP for estimation of the indigestibility from heat damage. 42 LITERATURE CITED Abdalla, H. 0., D. G. Fox, and R. R. Seaney. 1988. Variation in protein and fiber fractions in pasture during the grazing season. J. Anim. Sci. 66: 2663. Adamu, A. M., J. R. Russell, A. D. McGilliard, and A. Trenkle. 1989. Effects of added dietary urea on the utilization of maize stover silage by growing beef cattle. Anim. Feed Sci. Technol. 22: 227. Agricultural Research Council. 1984. Report of the Protein Group of the Agricultural Research Council Working Party on the Nutrient Requirements of Ruminants. Commonw. Agric. Bur., Slough, Engl. Aguilera, J. F., M. Bustos, and E. Molina. 1992. The degradability of legume seed meals in the rumen - effect of heat treatment. Anim. Feed Sci. Technol. 36: 101. Aharoni, Y. , H. Tagari, and R. C. Boston. 1991. A new approach to the quantitative estimation of nitrogen metabolic pathways in the rumen. Br. J. Nutr. 66: 407. Akin, D. E. 1979. Microscopic evaluation of forage digestion by rumen microorganisms - a review. J. Anim. Sci. 48: 701. Albrecht, K. A. and R. E. Muck. 1991. Proteolysis in ensiled forage legumes of varying tannin concentration. Page 76 in US Dairy Forage Research Center, 1991 Research Summaries, Madison, WI. Allen, M. S. and D. R. Mertens. 1988. Evaluating constraints on fiber digestion by rumen microbes. J. Nutr. 118: 261. Argos, P., K. Pedersen, M. D. Marks, and B. A. Larkins. 1982. A structural model for maize zein proteins. J. Biol. Chem. 257: 9984. Askerlund, P, C. Larsson, and S. W. Widell. 1989. Cytochromes of plant plasma membranes. Characterization by absorbance difference spectrophotometry and redox titration. Phyiologia Plantarum 76: 123. Association of Official Agricultural Chemists. 1975. Offical Methods of Analysis. 12th edition. A. O. A. C., Washington, D. C. - Assoumani, M. B., F. Vedeau, L. Jacquot, and C. J. Sniffen. 1992. Refinement of an enzymatic method for estimating the theoretical degradability of proteins in feedstuffs g for ruminants. Anim. Feed Sci. Technol. 39: 357. .. )7 I‘T‘IFH 43 Barrett, A. J. 1985. Chapter 1: The classes of proteolytic enzymes. Plant Proteolytic Enzymes. 1' 1 Barry, T. N. and T. W. Manley. 1984. The role of condensed tannins in the nutritional value of Lotus pendunculatus for sheep. Br. J. Nutr. 51: 493. Barry, T. N., A. Thompson, and D. G. Armstrong. 1977. Rumen fermentation studies on two contrasting diets. 1. Some characteristics of the in vivo fermentation, with special reference to the composition of the gas phase, oxidation/reduction state and volatile fatty acid proportions. J. Agric. Sci., Camb. 89: 183. Beever, D. E., S. B. Cammell, and A. Wallace. 1974. The digestion of fresh, frozen and dried perennial ryegrass. Proc. Nutr. Soc. 33: 73A. Bensadoun, A. and Weinstein. 1976. Assay of proteins in the presence of interfering materials. Anal. Biochem. 70: 241 Britton, R. A., T. J. Klopfenstein, R. Cleale, F. Goedekin, and V. Wilkerson. 1986. Methods of estimating heat damage in protein sources. Proc. Dist. Feed Conf. 41: 67. Brock, F. M., C. W. Forsberg, and J. G. Buchanan-Smith. 1982. Proteolytic activity of rumen microorganisms and effects of proteinase inhibitors. Appl. Environ. Microbiol. 44: 561. Broderick, G. A. 1987. Determination of protein degradation rates using a rumen in vitro system containing inhibitors of microbial nitrogen metabolism. Br. J. Nutr. 58: 463. Broderick, G. A. and D. R. Buxton. 1991. Genetic variation in alfalfa for ruminal protein degradability. Can. J. Plant Sci. 71: 755. Broderick, G. A. and M. K. Clayton. 1992. Rumen protein degradation rates estimated by non- linear regression analysis of Michaelis-Menten in vitro data. Br. J. N utr. 67: ‘ 27. 5 Broderick, G. A., J. H. Yang, and R. G. Koegel. 1993. Effect of steam heating alfalfa hay on utilization by lactating dairy cows. J. Dairy Sci. 76: 165. Buckhoz, H. F. andW. G. Bergen. 1973. Microbial phospholipid synthesisasamarker f for microbialprotein synthesisin the rumen. Appl. Microbial. 25: 504. ‘ Caton, J. 8., A. S. Freeman, and M. L. Galyean. 1988. Influence of protein . supplementation on forage intake,‘ in situ forage disappearance, ruminal fermentation ‘: 22d digesta passage rates in steers grazing dormant blue grama rangeland. J. Anim. Sci. 1 226 ‘ Cecava, M. J., N. R. Merchen, L. L. Berger, and G. C. Fahey, Jr. 1988. Effects of dietary energy level and protein source on site of digestion and duodenal nitrogen and . amino acid flows in steers. J. Anim. Sci. 66: 961 44 Chamberlain, D. G. and P. C. Thomas. 1979. Prospective laboratory methods for estimating the susceptibility of feed proteins to microbial breakdown in the rumen. Proc. Nutr. Soc. 38: 138A. Chen, G. and J. B. Russell. 1989. More monensin-sensitive, ammonia-producing bacteria from the rumen. Appl. Environ. Microbiol. 55: 1052. Chen, G., C. J. Sniffen, and J. B. Russell. 1987. Concentration and estimated flow of peptides from the rumen of dairy cattle: Effects of protein quantity, protein solubility, and feeding frequency. J. Dairy Sci. 70: 983. Ciszuk P. and J. E. Lindberg. 1988. Responses in feed intake, digestibility and nitrogen retention in lactating dairy goats fed increasing amounts of urea and fish meal. Acta Agric. Scand. 38: 381. Coenen, D. J. and A. Trenkle. 1989. Comparisons of expeller processed-and solvent- extracted soybean meals as protein supplements for cattle. J. Anim. Sci. 67: 565. Coto, G., R. S. Herrera, M. Perez, J. Alert, and S. Poppe. 1989. A study on the soluble and insoluble nitrogen fractions in star grass fertilized with high nitrogen doses. Cuban, J. Agric. Sci. 23: 115. Crawford, R. J. Jr., W. H. Hoover, C. J. Sniffen, and B. A. Crooker. 1978. Degradation of feedstuff nitrogen in the rumen vs. nitrogen solubility in three solvents. J. Anim. Sci. 46: 1768. Crooker, B. A., C. J. Sniffen, W. H. Hoover, and L. L. Johnson. 1978. Solvents for soluble nitrogen measurements in feedstuffs. J. Dairy Sci. 4: 437. Cruickshank, G. J., D. P. Poppi, and A. R. Sykes. 1989. Theoretical considerations in the estimation of rumen fractional outflow rate from various sampling sites in the digestive tract. Br. J. Nutr. 62: 229. Czerkawski, J. W. and C. Foulds. 1974. Methods for determining 2-6-diaminopimelic acid and 2-aminoethylphosphoric acid in gut contents. J. Sci. Food Agric. 25: 45. Dahlberg, E. M., M. D, Stern, and F. R. Ehle. 1988. Effects of forage source on ruminal microbial nitrogen metabolism and carbohydrate digestion in continuous culture. J. Anim. Sci. 66: 2071. Dhanoa, M. S. 1988. Research note on the analysis of dacron bag data for low degradability feeds. Gr. Forage Sci. 43: 441. Duvick, D. N. 1961. Protein granules of maize endosperm cells. Cereal Chem. 38: 374. Esen, A. 1980. Estimation of protein quality and quantity in corn (Zea mays L.) by assaying protein in two solubility fractions. J. Agric. Food Chem. 28: 529. Fahmy, W. G., J. Aufrere, D. Graviou, C. Demarquilly, and K. El-Shazly. 1991. Comparison between the mechanism of protein degradation of 2 cereals by enzymatic and in situ methods, using gel electrophoresis. Anim. Feed Sci. Technol. 35: 115. :l'Tl warn KEITH 45 Faichney, G. J. 1975. The use of markers to partition digestion within the gastro- intestinal tract of ruminants. Pages 277-291 in Digestion and Metabolism in the Ruminant. ed. 1. W. McDonald and A.C.I Warner, University of New England, Arrnidale. Faichney, G. J. 1980. The use of markers to measure digesta flow from the stomach of sheep fed once daily. J. Agric. Sci., Camb. 94: 313. Faichney, G. J. 1992. Application of the double-marker method for measuring digesta kinetics to rumen sampling in sheep following a dose of the markers or the end of their continuous infusion. Aust. J. Agric. Res. 43: 277. Ferguson, J. D. and W. Chalupa. 1989. Symposium: interactions of nutrition and reproduction. Impact of protein nutrition on reproduction in dairy cows. J. Dairy Sci. : 746. Fluharty, F. L. and S. C. Loerch. 1989. Chemical treatment of ground corn to limit ruminal digestion. Can. J. Anim. Sci. 69: 173. Folman, Y., H. Neumark, M. Kaim, and W. Kaufmann. 1981. Performance, rumen and blood metabolites in high-yielding cows fed varying protein percents and protected soybean. J. Dairy Sci. 64: 759. Fuchigami, M., T. Senshu, and M. Horiguchi. 1989. A simple continuous culture system for rumen microbial digestion study and effects of defaunation and dilution rates. J. Dairy Sci. 72: 3070. Fujimaki, T., Y. Kobayashi, M. Wakita, and S. Hoshino. 1992. Amino acid supplements -— a least combination that increases microbial yields of washed cell suspension from goat rumen. J. Anim. Physiol. Anim. Nutr. -- Zeitschrift Fur Tierphysiologie Tieremahrung und Futtermittelkunde 67: 41. ‘ Furchtenicht, J. E. and G. A. Broderick. 1987. Effect of inoculum preparation and gietary energy on microbial numbers and rumen protein degradation activity. J. Dairy ci. 70: 1404. Furchtenicht, J. E. and G. A. Broderick. 1987. Effect of inoculum preparation and gietary energy on microbial numbers and rumen protein degradation activity. J. Dairy ci. 70: 1404. Ganev, G., E. R. Orskov, and R. Smart. 1979. The effect of roughage or concentrate feeding and rumen retention time on total degradation of protein in the rumen. J. Agric. 801., Camb. 93: 651. Gasa, J., K. Holtenius, J. D. Sutton, M. S. Dhanoa, and D. J. Napper. 1991. Rumen fill and digesta kinetics in lactating friesian cows given 2 levels of concentrates with 2 types of grass silage ad—lib. Br. J. Nutr. 66: 381. Gianazza, E., V. Viglienghi, P. G. Righetti, F. Salamini, and C. Soave. 1977. Amino acrd composition of zein molecular components. Phytochem. 16: 315. 46 Goering, H. K. 1972. A laboratory assessment on the frequency of over-heating in commercial dehydrated alfalfa samples. J. Anim. Sci. 43: 869. Goering, H. K., C. H. Gordon, R. W. Hemken, D. R. Waldo, P. J. Van Soest, and L. W. Smith. 1972. Analytical estimates of nitrogen digestibility in heat damaged forages. J. Dairy Sci. 55: 1275. Gottschalk, G. 1986. Bacterial Metabolism. 2nd ed. Springer-Verlag, New York. Grabber, J. H. and G. A. Jung. 1991. In viu'o disappearance of carbohydrates, phenolic acids, and lignin from parenchyma and sclerenchyma cell walls isolated from cocksfoot. J. Sci. Fd. Agric. 57: 315. Griffin, T. 8., 0. B. Hesterrnan, and S. R. Rust. 1991. Alfalfa maturity and cultivar effects on chemical and in situ estimates of protein degradability. Page 23 in American Forage and Grassland Council, Interpretive Summaries. Amway Grand Plaza Hotel, Grand Rapids, MI. Gunn, K. J. and N. P. McMeniman. 1989. Incorporation of 358 into rumen microbial protein. Proc Nutr Soc (Aust) 14: 141. Hach, C. H., S. V. Brayton, and A. B. Kopelove. 1985. A powerful kjeldahl nitrogen method using peroxymonosulfuric acid. J. Agric. Food Chem. 33: 1117. Hagerrnan, A. E., C. T. Robbins, Y. Weerasuriya, T. C. Wilson, and C. Mcarthur. 1992. Tannin chemistry in relation to digestion. J. Range Manag. 45: 57. Halverson, A. W., G. D. Williams, and G. D. Paulson. 1968. Aspects of sulfate utilization by the microorganisms of the ovine rumen. J. Nutr. 95: 363. Hamilton, B. A., J. R. Ashes, and A. W. Carmichael. 1992. Effect of formaldehyde- treatedasgnflower meal on the milk production of grazing dairy cows. Austr. J. Agric. Res. 4 : 79. Hanley, T. A., C. T. Robbins, A. E. Hagerman, and C. Mcarthur. 1992. Predicting digestible protein and digestible dry matter in tannin-containing forages consumed by ruminants. Ecology 73:537. Hartley, R. D. and D. E. Akin. 1989. Effect of forage cell wall phenolic acids and derivatives on rumen micorflora. J. Sci. Food Agric. 49: 405. Harvey, B. M. R. and A. Oaks. 1974. The hydrolysis of endosperm protein in Zea mays. Plant Physiol. 53: 453. Hazlewood, G. P.,C. G. Orpin, Y. Greenwood, M. E. Black. 1983. Isolation of proteolytic rumen bacteria by use of selective medium containing leaf fraction 1 protein (rubulosebisphosphate carboxylase). Appl. Environ. Microbiol. 45: 17 80. Henderickx, H. K., D. I. Demeyer, C. J. Van Nevel. 1972. Problems in estimating microbial protein synthesis in the rumen. Pages 57-68 in Tracer Studies on Non-Protein Nitrogen for Ruminants. Inter. Atomic Energy Agency. Vienna. 47 Hoffman, P. C., S. J. Sievert, R. D. Shaver, and D. K. Combs. 1992. In situ CP degradation kinetics of perennial forages. J. Dairy Sci. 75(Suppl. 1):210. Hunter, R. A. and B. D. Siebert. 1987. The effect of supplements of rumen-degradable protein and formaldehyde-treated casein on the intake of low—nitrogen roughages by Bos taurus and Bos indicus steers at diferent stages of maturity. Aust. J. Agric. Res, 38: 209. Jancarik, A. and M. Proksova. 1970. The breakdown of protein in the rumen in relation to the physical and chemical characters of the rumen juice. Pages 304-306 in Nutr. Proc. 8th Intern. Congr. Prague. Excerpta Medica, Amsterdam. Janicki, F. J., and C. C. Stallings. 1988. Degradation of crude protein in forages determined by 1n vitro and' in situ procedures. J. Dairy Sci. 71: 2440 Javorsky, P., I. Havassy, E. Rybosova, M. Kralova, K. Horsky, and K. Kosta. 1986. Synthesis of bacterial nitrogen substances in the sheep rumen from nitrogen-15 labelled urea. J. Anim. Physiol. Anim. Nutr. 5: 281. Johnson, D. E., G. M. Ward, and J. Torrent. 1992. The environmental impact of bovine somatotropin use in dairy cattle. J. Environ. Quality 21: 157. Jones, W. T. and J. L. Mangan. 1977. Complexes of the condensed tannins of sainfoin (Onobrychis viciifolia Scop) with fraction 1 leaf protein and with sub-maxillary mucoprotein, and their reversal by polyethylene glycol and pH. J. Sci. Food Agric. 28: 126. Kephart, K. D. and A. Boe. 1991.Rumen degradable protien of Rumen degradable protein of cool- and warm—season forage grasses. Page 24 in Amer. Forage and Grassland Council, Interpretive Summaries. Amway Grand Plaza Hotel, Grand Rapids, MI. 3 Kopecny, J. and R. J. Wallace. 1982. Cellular location and some properties of proteolytic enzymes of rumen bacteria. Appl. Environ. Microbiol. 43: 1026. Krishnamoorthy, U, T. V. Muscato, C. J. Sniffené andP. J. Van Soest 1982. Nitrogen fractions 1n selected feedstuffs. J. Dairy Sci. 65:21.7 Krishnamoorthy, U., C. J. Sniffen, M. D. Stern, and P. J. Van Soest. 1983. Evaluation of a mathematical model of rumen digestion and an in vitro simulation of rumen proteolysis to estimate the rumen-undegraded nitrogen content of feedstuffs. Br. J. Nutr. 50: 555. Kristensen, E. S., P. D. Moller, and T. Hvelplund. 1982. Estimation of the effective protein degradability 1n the rumen of cows using the nylon bag technique combined with the outflow rate. Acta Agric. Scand. 32: 1.23 Kung, L. Jr. and J. T. Huber. 1983. Performance of high producing cows in early 12actation fed protein of varying amounts, sources, and degradability. J. Dairy Sci. 66: 27 48 Larnport, D.T.A. 1980. Structure and function of plant glycoproteins. Pages 501—541 in The Biochemistry of Plants, vol. 3, Academic Press Inc., New York. Lee, K. B., D. Loganathan, Z. M. Merchant, and R. J. Linhardt. 1990. Carbohydrate analysis of glycoproteins: a review. Appl. Biochem. Biotech. 23: 53. Lehninger, A. L. 1975. Biochemistry. The Molecular Basis of Cell Structure and Function. 2nd edition. Worth Publishers, Inc. New York. Lindberg, J. E. 1985. Estimation of rumen degradability of feed proteins with the in sacco technique and various in vitro methods. A review. Acta Agric. Scand. Suppl. 25: 64. Lindberg, J. E. 1988. Influence of cutting time and N fertilization of the nutritive value of timothy: 2. Estimates of rumen degradability of nitrogenous compounds. Swedish J. Agric. Res. 18: 85. Lindberg, J. E., H. Bristav, and A. R. Manyenga. 1989. Excretion of purines in the urine of sheep in relation to duodenal flow of microbial protein. Swedish J. Agric. 19: 45. Lindberg, J. E. and K. G. J acobsson. 1990. Nitrogen and purine metabolism at varying energy and protein supplies in sheep sustained on intragastric infusion. Br. J. Nutr. 64: 359. Lindberg, J. E. and I. Olsson. 1983. Live weight gain in intensively reared bulls fed rations with low and high degradable protein. Pages 243-246 in IVth Int Symp. Protein metabolism and nutrition, Clermont—Ferrand (France) Sept. 5-9. NRA Publ. 1983. II (les colloques de INRA no 16), Clermont, France. 1 Ling, J. R. and Buttery, P. J. 1978. The simultaneous use of ribonucleic acid, 358, 2,6,- ‘ diaminopimelic acid and 2-aminoethylphosphonic acid as markers of microbial nitrogen entering the duoddenum of sheep. Br. J. Nutr. 39: 165. Lowry, O. H., N. O. Rosenbrough, A. L. Fan, and R. J. Randall. 1951. Protein measurement with the Folin-phenol reagent. J. Biol. Chem. 193: 265. Lyttleton, J. W. 1973. Proteins and nucleic acids. Pages 63-103 in Chemistry and Biochemistry of Herbage. ed. G. W. Butter and R. W. Baily. Academic Press, New York, NY. Madsen, J. 1985. Protein evaluation for ruminants. Acta Agric. Scand. Suppl. 25: 9. Madsen, J. and T. Hvelplund. 1988. The influence of different protein supply and feeding level on pH, ammonia concentration and microbial protein synthesis in the ‘ rumen of cows. Acta Agric. Scand. 38: 115. Mahadevan, 8., J. D. Erfle, and F. D. Sauer. 1979. A colorimetric method for the determination of proteolytic degradation of feed proteins by rumen microorganisms. J. Anim. Sci. 48: 947. 49 Mahadevan, 8., J. D. Erfle, and F. D. Sauer. 1980. Degradation of soluble and insoluble proteins by Bacteroides amylophilus protease and by rumen microorganisms. J. Anim. Sci. 50: 723. Mahadevan, 8., F. D. Sauer, and J. D. Erfle. 1987. Preparation of protease from mixed rumen microorganism and its use for the in vitro determination of the degradability of true protein in feedstuffs. Can. J. Anim. Sci. 67: 55. Mangan, J. L. 1982. 2.1 The nitrogenous constituents of fresh forages. Pages 25-39 in Forage Protein in Ruminant Anim. Prod., ed. Thomson, D. J. et al., Occasional Publication No. 6, Br. Soc. Anim. Prod., London. Manson, F. J. and J. D. Leaver. 1988. The influence of dietary protein intake and of hoof trimming on lameness in dairy cattle. Anim. Prod. 47: 191. Marder, J. B. and J. Barber. 1989. The molecular anatomy and function of thylakoid proteins. Plant, Cell and Environment 12: 595. Marinucci, M. T., B. A. Dehority, and S. C. Loerch. 1992. In vitro and in vivo studies of factors affecting digestion of feeds in synthetic fiber bags. J. Anim. Sci. 70: 296. Mathers, J. C. and E. L Miller. 1980 A simple procedure using 358 incorporation for the measurement of microbial and undegraded food plotein in ruminant digesta. Br. J. Nutr. 43. 503. Mathison, G. W. and L. P. Milligan. 1971. Nitrogen metabolism in sheep. Br. J. Nutr. 25: 351. Matin, A. 1992. Physiology, molecular biology and applications of the bacterial ? starvation response. Journal of Applied Bacteriology 73: S49. 1 Matras, J. and R. L. Preston. 1989. The role of glucose infusion on the metabolism of nitrogen in ruminants. J. Anim. Sci. 67: 1642. McAllan, A. B. and E. S. Griffith. 1987. The effects of different sources of nitrogen supplementation on the digestion of fibre components in the rumen of steers. Anim. Feed Sci. Technol. 17:65. McAllan, A. B. and R. H. Smith. 1983A. Estimation of flows of organic matter and nitrogen components in postruminal digesta and effects of level of dietary intake and physical form of protein supplement on such estimates. Br. J. Nutr. 49: 119. McAllan, A. B. and R. H. Smith. 1983B. Influence of different dietary nitrogen sources on fibre digestion in the rumen. Pages 255-257 in IVth Int. Symp. Protein Metabolism and Nutrition, Clermont-Ferrand (France) Sept. 5-9. INRA Publ. 1983. II (les colloques de INRA no 16) McDonald, 1. 1981. A revised model for the estimation of protein degradability in the rumen. J. Agric. Sci., Camb. 96: 251. 50 McDonald, P. 1982. The effect of conservation processes on the nitrogenous components of forages. Pages 41-49 in Forage Protein in Ruminant Animal Prod. Occasional Publication No. 6 Br. Soc. of Anim. Prod. McKersie, B. D. 1981. Proteinases and peptidases of alfalfa herbage. Can. J. Plant Sci. 61: 53. McKersie, B. D. 1985. Effect of pH on proteolysis in ensiled legume forage. Agron. J. 77: 80. Mehrez, A. Z. and E. R. Qrskov. 1977. A study of the artificial fibre bag technique for determining the digestibility of feeds 1n the rumen. J. Agric. Sci., Camb. 88: 64 5. Meissner, H. H. and U. Todtenhofer. 1989. Influence of casein and glucose or starch supplementation in the rumen or abomasum on utilization of Eragrostis curvula hay by sheep. S. Afr. J. Anim. Sci. 19: 43. Mercer, J. R., S. A. Allen, and E. L. Miller. 1980. Rumen bacterial protein synthesis and the proportion of dietary protein escaping degradation in the rumen of sheep. Br. J. Nutr. 43: 421. Messman, M. A. and W. P. Weiss. 1992. Extraction and colorimetric determination of forage protein. J. Dairy Sci. 75(Suppl. 1):l75. Messman, M. A., W. P. Weiss, and D. O. Erickson. 1992A. Effects of nitrogen fertilization and maturity of bromegrass on nitrogen and amino acid utilization by cows. J. Anim. Sci. 70: 566. Messman, M A., W. P. Weiss, and M. E. Koch. 1992B. Electrophoresis as a tool to ascertain effects of preservation method on forage protein composition. Pages 35 to 39 \in Ohio Agric. Experiment Station Research Highlights, Wooster, OH. Meyer, J. H. F. andR. I. Mackie. 1986. Microbiologicalevaluation of the 1ntrarum1nal in sacculus d1gest1ontechn1que Appl. Environ. Mic1obiol. 51. 622 Meyer, J. H. F. S. I. van der Walt, and H. M. Schwartz. 1986. The influence of diet and protozoa] numbers on the breakdown and synthesis of protein in the rumen of sheep. J. Anim. Sci. 62: 509. Michalet-Doreau, B. and M. Y. Ouldbah. 1992. In vitro and in sacco methods for the estimation of dietary nitrogen degradability in the rumen - a review. Anim. Feed Sci. Technol. 40: 57. Mohamed, O. E. and R. H. Smith. 1977. Measurement of protein degradation in the rumen. Proc. Nutr. Soc. Soc. 36:152A. Mosimanyana, B. M. andD. N. Mowat. 1992. Rumen protection of heat- treated soybean proteins. Can. J. Anim. Sci. 72: Mullahey, J. J, S. S. Waller, K. J. Moore, L. E. Moser, andT. J. Klopfenstein.1992.1n iiéu ruminal protein degradation of switchgrass and smooth bromegrass. Agron. J. 84: 3. 51 Murphy, J. J. and J. J. Kennelly. 1987. Effect of protein concentration and protein source on the degradability of dry matter and protein in situ. J. Dairy Sci. 70: 1841. Nader, C. J. and D. J. Walker. 1970. Metabolic fate of cysteine and methionine in rumen digesta. Appl. Microbiol. 20: 677. National Research Council. 1985. Ruminant Nitrogen Usage. Subcommittee on Nitrogen Usage in Ruminants Committee on Animal Nutrition Board on Agriculture. National Academy Press. Washington, D. C. National Research Council. 1989. Nutrient Requirements of Dairy Cattle. 6th rev. ed. Natl. Acad. Sci., Washington, DC. Neutze, S. A., R. L. Smith, and W. A. Forbes. 1993. Application of an inhibitor in vitro method for estimating rumen degradation of feed protein. Anim. Feed Sci. Technol. 40: 251. Newbold, J. R., P. C. Gamsworthy, P. J. Buttery, D. J. A. Cole, and W. Haresign. 1987. Protein nutrition of growing cattle: Food intake and growth responses to rumen degradable protein and undegradable protein. Anim. Prod. 45: 383. Nia, S. A. M. and J. R. Ingalls. 1992. Effect of heating on canola meal protein degradation in the rumen and digestion in the lower gastrointestinal tract of steers. Can. J. Anim. Sci. Sci.72: 83. Nocek, J. E., K. A. Cummins, and C. E. Polan. 1979. Ruminal disappearance of crude groteisn7 and dry matter in feeds and combined effects 1n formulated rations. J. Dairy Sci. 2 Nocek, J. E., J. H. Herbein, and C. E. Polan. 1983. Total amino acid release rates of soluble and insoluble protein fractions and concentrate feedstuffs by Streptomyces griseus. J. Dairy Sci. 66: 1663. Nolan, J. V., W. B. Norton, and R. A. Leng. 1972. Dynamic aspects of nitrogen metabolism in sheep. Pages 13-24 in Tracer Studies on Non-Protein Nitrogen for Ruminants. Inter. Atomic Energy Agency. Vienna. Nugent, J. H., andJ. L. Mangan. 1981. Characteristics of the rumen proteolysis of fraction 1 (188) leaf protein from luceme (Medicago sativa L). Br. J. Nutr. 46: 39. L Ohshirna, M. and P. McDonald. 1978. A review of the changes in nitrogenous . compounds of herbage during ensilage. J. Food Agric. 29:497. Oke, B. O. and S. C. Loerch. 1992. Effects of energy and amino acid supply to the small intestine on amino acid metabolism. J Nutr. Biochem. 3. 63. @rskov, E. R. M. Hughes- J-ones, and I McDonald. 1980. Degradability of protein supplements and utilization of undegraded protein by high— producing dairy cows. Pages {5- -98 1n Recent Advances 1n Animal Nutrition. ed. W. Haresign, Butterworths, ondon. 52 firskov, E. R., and I. McDonald. 1979. The estimation of protein degradability in the rumen from incubation measurements weighted according to rate of passage. J. Agric. Sci., Camb. 92: 499. Ourry, A., B. GonzaleLand J. Boucaud. 1989. Osmoregulation and role of nitrate during regrowth after cutting of ryegrass (Lolium perenne). Physiologia Plantarum 76: 177. Papadopoulos, M. C. 1989. Effect of processing on high-protein feedstuffs: a review. Biol. Wastes 29: 123. Paulis, J. W. 1981. Disulfide structures of zein proteins from corn endosperm. Cereal Chem. 58: 542. Petit, H. V. and G. F. Tremblay. 1992. In situ degradability of fresh grass and grass conserved under different harvesting methods. J. Dairy Sci. 75: 774. Pichard, G. and P. J. Van Soest. 1977. Protein solubility of ruminant feeds. Page 91 in Cornell Nutr. Conf. Feed Manufacturers, Syracuse, NY. Polan, C. E. 1988. Update: dietary protein and microbial protein contribution. J. of Nutr. 118: 242. Poos-Floyd, M., T. Klopfenstein, and R. A. Britton. 1985. Evaluation of laboratory techniques for predicting ruminal protein degradation. J. Dairy Sci. 68: 829. Puchala. R. and G. W. Kulasek. 1992. Estimation of microbial protein flow from the rumen of sheep using microbial nucleic acid and urinary excretion of purine derivatives. Canadian J. Anim. Sci. 72:821. Raab, L, B. Cafantaris, T. Jilg, and K. H. Menke. 1983. Rumen protein degradation and biosynthesis l. a new nethod for determination of protein degradation in rumen fluid in vitro. Br. J. Nutr. 50: 569. Robinson, P. H., and J. J. Kennelly. 1988. Influence of intake of rumen undegradable protein on milk production of late lactation holstein cows. J. Dairy Sci. 71: 2135. Rooke, J. A. and D. G. Armstrong. 1989. The importance of the form of nitrogen on microbial protein synthesis in the rumen of cattle receiving grass silage and continuous intrarumen infusions of sucrose. Br. J. Nutr. 61: 113. Rooke, J. A., H. A. Greife, and D. G. Armstrong. 1984. The effect of in sacco rumen incubation of a grass silage upon the total and D-amino acid composition of the residual silage dry matter. J. Agric. Sci., Camb. 102: 695. Russell, J. B., C. J. Sniffen, and P. J. Van Soest. 1983. Effect of carbohydrate limitation on degradation and utilization of casein by mixed rumen bacteria. J. Dairy Sci. 66: 763. Ruxton, I. B. and P. McDonald. 1974. The influence of oxygen on ensilage. 1. Laboratory studies. J. Sci. Food Agric. 12: 706. 53 Salobir, K., A. Pen, and O. Muck. 1970. The importance of solubility of proteins in feed for the evaluation of their propriety for ruminants. Page 329 in Proc. 1st Yugoslav Intern. Conf. Anim. Prod. Sanderson, M. A. and W. F. Wedin. 1989A. Nitrogen concentrations in cell wall and lignocellulose of smooth brome grass. Grass Forage Sci. 44: 151. Sanderson, M. A. and W. F. Wedin. 1989B. Nitrogen in the detergent fibre fractions of temperate legumes and grasses. Grass Forage Sci. 44: 159. Santos, K. A., M. D. Stern, and L. D. Satter. 1984. Protein degradation in the rumen and amino acid absorption in the small intestine of lactating dairy cattle fed various protein sources. J. Anim. Sci. 58: 244. Sathe, S. K., A. C. Mason, C. M. Weaver. 1989. Thermal aggregation of soybean (glycine max L.) sulfur-rich protein. J. Food Sci. 54: 319. Schmitter, B. M. and T. Rihs. 1989. Evaluation of a macrocombustion method for total nitrogen determination in feedstuffs. J. Agric. Food Chem. 37: 992. Schwab, C. G., c. K. Bozak, N. L. Whitehouse, and M.M.A. Mesbah. 1992. Amino acid limitation and flow to the duodenum at four stages of lactation. 1. Sequence of lysine and methionine limitation. J. Dairy Sci. 75: 3486. Segel, I. H. 1976. Biochemical Calculations. 2nd ed., John Wiley and Sons, Inc., New York. Siddons, R. C., D. E. Beever, J. V. Nolan, A. B. McAllan, and J. C. Macrae. 1979. :Estimation of microbial protein in duadenal digesta. Ann. Rech. Vet. 10: 286. Sklan, D. 1989. In vitro and in vivo rumen protection of proteins coated with calcium soaps of long-chain fatty acids. J. Agric. Sci., Camb. 112: 79. Smith, C. R. J r., F. R. Earle, I. A. Wolff, and Q. Jones. 1959. Comparisons of solubility characteristics of selected seed proteins. J. Agric. Food Chem. 7: 133. Smith, R. H., A. B. McAllan, D. Hewitt, and P. E. Lewis. 1978. Estimation of amounts of microbial and dietary nitrogen compounds entering the duodenum of cattle. J. Agric. Sci., Camb. 90: 557. Spencer, D., T. J. V. Higgins, M. Freer, H. Dove, and J. B. Coombe. 1988. Monitoring the fate of dietary proteins in rumen fluid using gel electrophoresis. Br. J. Nutr. 60: 241. Staples, C. R. and D. S. Lough. 1989. Efficacy of supplemental dietary neutralizing agents for lactating dairy cows. A review. Anim. Feed Sci. Technol. 23: 277. Stern, M. D. and L. D. Satter. 1984. Evaluation of nitrogen solubility and the dacron bag technique as methods for estimating protein degradation in the rumen. J. Anim. Sci. 58: 7 14. 54 Teeter, R. G. and F. N. Owens. 1983. Characteristics of water soluble markers for measuring rumen liquid volume and dilution rate. J. Anim. Sci. 56: 717 Terramoccia, S., S. Puppo, L. Rizzi, and F. Martillotti. 1992. Comparison between in- sacco and in vitro protein rumen degradability. Annales de Zootechnie 41: 20. Thomsen, K. V. 1985. The specific nitrogen requirements of rumen microorganisms. Acta Agric. Scand. Suppl. 25: 125. Tyrrell, HE, P. W. Moe, and W. P. Flatt. 1970. Influence of excess protein intake on energy metabolism of the dairy cow. Pages 68-71 in Publ. 16, European Assoc. Anim. Prod., London. Van Bruchem, J., A. K. Kies, R. Bremmers, M. W. Boxch, H. Boer, and P. W. M. Van Adrichem. 1988. Digestion of alfalfa and grass silages in sheep. 2. Digestion of protein in reticulorumen and intestine. Neth. J. Agric. Sci. 36: 365. Van Ramshorst, H. and P. C. Thomas. 1988. Digestion in sheep of diets containing barley chemically treated to reduce its ruminal degradability. J. Sci. Food Agric. 42: 1. Van Soest, P. J. 1965. Use of detergents in analysis of fibrous feeds. III. Study of effects of heating and drying on yield of fiber and lignin in forages. J. Assoc. Offic. Anal. Chem. 48: 785. Van Soest, P. J. 1982. Chapter 4. Fecal composition, mathematics of digestion balances and markers. Page 39 in Nutritional Ecology of the Ruminant, Cornell University Press, Ithaca, NY. Varvikko, T. and J. E. Lindberg. 1985. Estimation of microbial nitrogen in nylon-bag residues by feed 15-N dilution. Br. J. Nutr. 54: 473. .Veen, W. A. G, J. Veling, and Y. T. Bakker. 1988. The influence of slowly and rapidly degradable concentrate protein on nitrogen balance, rumen fermentation pattern and blood parameters in dairy cattle. Nether. J. of Agric. Sci. 36. 127. Vik-Mo, L. 1989. Degradability of forages in sacco 2. Silages of grasses and red clover at two cutting times, with formic acid and without additive. Acta Agric. Scand. 39: 53 Visek, W. J. 1984. Ammonia: its effects on biological systems, metabolic hormones, and reproduction. J. Dairy Sci. 67: 481. Voss V. L, D. Stehr, L. D. Satter, andG. A. Broderick. 1988. Feeding lactating dairy cows proteins resistant to ruminal degradation. J. Dairy Sci. 71: 2428. "Waghom, G. C. 1991. Electronegativity and redox potential of rumen digesta in situ in ,cows eating fresh luceme. N. Z. J. Agric. Res. 34: 359 ‘Wagner, F. W. 1985. Chapter2. Assessment of methodology for the purification, characterization, and measurement of proteases. Pages 17- 39 1n Plant Proteolyuc Enzymes. 1: l7. 55 Waldo, D. R. and H. K. Goering. 1979. Insolubility of proteins in ruminant feeds by four methods. J. Anim. Sci. 49: 1560. Walker, D. J. and C. J. Nader. 1968. Method for measuring microbial growth in rumen contents. Appl. Microbiol. 16: 1124. Walker, D. J. and C. J. Nader. 1975. Measurement in vivo of rumen microbial protein synthesis. Aust. J. Agric. Res. 26: 689. wallace, R. J. 1992. Gel filtration studies of peptide metabolism by rumen microorganisms. J. Sci. Fd. Agric. 58: 177. Wallace, R. J ., G. A. Broderick, and M. A. Brammall. 1987. Protein degradation by ruminal microorganisms from sheep fed dietary supplements of urea, casein, or albumin. Appl. and Environ. Microbiol. 53: 751. Wallace, R. J. and M. L. Falconer. 1992. In vitro studies of conditions required to protect protein from ruminal degradation by heating in the presence of sugars. Anim. Feed Sci. Technol. 37: 129. Wallace, R. J. and J. Kopecny. 1983. Breakdown of diazotized proteins and synthetic substrates by rumen bacterial proteases. Appl. Environ. Microbiol. 45: 212. Warner, A. C. I. and B. D. Stacy. 1968. The fate of water in the rumen 1. A critical appraisal of the use of soluble markers. Br. J. Nutr. 22: 369. Waters, C. J, M. A. Kitcherside, and A. J. F. Webster. 1992. Problems associated with estimating the digestibility of undergraded dietary nitrogen from acid- -detergent insoluble nitrogen. Anim. Feed Sci. Technol. 39: 279. ‘ Watkins, K. L., T. L. Veum, and G. F. Krause. 1987. Total nitrogen determination of various sample types: a comparison of the Hach, Kjeltec, and Kjeldahl methods. J. Assoc. Offic. Anal. Chem. 70: 410. Weiss, W. P., H. R. Conrad, andW. L. Shockey. 1986A. Amino acid profiles of heat- damaged grasses. J. Dairy Sci. 69: 1824. Weiss, W. P. ,H. R. Conrad, andW. L. Shockey. 1986B. Digestibility of n1trogen 1n heat-damaged alfalfa. J. Dairy Sci. 69: 2658. , Weiss, W. P, D. O. Erickson, G. M. Erickson, and G. R. Fisher. 1989. Barley distillers grains as a protein supplement for dairy cows. J. Dairy Sci. 72: 980. Whitelaw, F. G., J. M. Eadie, L. A. Bruce, and W. J. Shand. 1984. Microbial protein synthesis in cattle given roughage concentrate and all concentrate diets: the use of 2,6- diaminopimelic acid, 2-aminoethylphosphonic acid and 35S as markers. Br. J. Nutr. 52: 249. Wilson, J. R., D. E. Akin, M. N. McLeod, and D. J. Minson. 1989. Particle size reduction of the leaves of a tropical and a temperate grass by cattle. II. Relation of anatomical structure to the process of leaf breakdown through chewing and digestion. Grass Forage Sci. 44: 65. 56 Windham, W. R. 1989. Potential characterization of a protein-carbohydrate complex from the rumen of steers fed high quality forages. J. Agric. Food Chem. 37: 912 Wohlt, J. E., C. J. Sniffen, and W. H. Hoover. 1973. Measurement of protein solubility in common feedstuffs. J. Dairy Sci. 56: 1052. Wong, E. 1973. Plant Phenolics. Pages 265-322 in Chemistry and Biochemistry of Herbage. ed. G. W. Butler and R. W. Bailey. Academic Press, New York, NY. Zinn, R. A. and F. N. Owens. 1986. A rapid procedure for purine measurement and its use for estimating net ruminal protein synthesis. Can. J. Anim. Sci. 66: 157. Zorrilla-Rios, J. and G. W. Horn. 1989. Effect of ammoniation and energy supplementation on the utilization of wheat straw by sheep. Anim. Feed Sci. Technol. 22: 305. ef ad ov ga aff Prc 57 Reprinted with permission from J. Sci. Food Agric. (Kohn. R. A. and M. S. Allen. 1992. 58:215-220. CHAPTER 2: STORAGE OF FRESH AND ENSILED FORAGFS BY FREEZING AFFECTS FIBRE AND CRUDE PROTEIN FRACTIONS.I Abstract The effect of freezing on fibre and crude protein fractions of forages was determined. Fresh and ensiled luceme and fresh bromegrass were processed immediately after collection or stored at -25 °C for l d, 1 mo, 6 mo or 12 mo before drying at 55 °C. Samples were frozen quickly by submersion in liquid nitrogen or slowly at -25 °C. Samples which were not frozen were processed immediately or after 1 h delay at room temperature. All treatments were replicated (n=3). Samples were analyzed for crude protein (CP), trichloroacetic acid soluble CP (TSCP), phosphate buffer soluble CP (BSCP), neutral detergent insoluble CP (NDCP), acid detergent insoluble CP (ADCP), neutral detergent fibre (NDF), acid detergent fibre (ADF), acid detergent sulfuric acid lignin (lignin) and ash. Freezing decreased BSCP and increased NDCP by more than 40% for bromegrass. Freezing also changed NDF, ADF, lignin, ash, CP and ADCP in different ways depending on forage type and length of time frozen. No significant effects were observed for method of freezing or a l h delay in processing. An additional experiment showed that freeze-drying resulted in less insoluble protein than oven-drying. Prior freezing of forages appeared inconsistently to change the extent of gaseous loss during drying, and resulted in precipitation of protein. These changes also affected fibre estimates. Forages should not be frozen and thawed before analysis of protein or fibre fractions. 'AbsmwesenwdameanwfingofmeAmdmnDaindweeAssodmimu-fllum l990.Raleigh,NC,USA. Frb Ru H0 fib 58 Key words Fibre, NDF, ADF, protein fractions, freezing, freeze-drying, forage analysis. 1. Introduction Ruminants normally consume fresh forages and silages without prior freezing. However, freezing is a common method of storing forage samples prior to analysis of fibre and nitrogen fractions. Neutral detergent fibre (NDF) is related to forage intake by ruminants (Van Soest 1965a; Mertens 1973) and is used to measure fibre requirements of dairy cattle (Mertens 1987). Acid detergent fibre (ADF) is negatively related to digestible energy content of ruminant feeds (Van Soest 1965a). Lignin is an indicator of feed quality and is associated with decreased digestibility of feeds (Woodman and Stewart 1932; Jung 1989). Crude protein (CP) soluble in trichloroacetic acid (TCA), neutral buffers, or neutral detergents generally is more rapidly converted to ammonia by ruminal microbes than insoluble protein (Mahadevan etal 1980; Nocek etal 1983; Van Soest and Sniffen 1984; Lindberg 1988). The rate of protein degradation may, in turn, affect protein utilization by ruminants (National Research Council, 1985). Crude protein which is insoluble in acid detergent (ADCP) is a measure of heat damage to silage or hay that renders the carbohydrate and protein indigestible (Van Soest 1965b; Weiss etal 1986). Freezing of vegetative material and foods is associated with precipitation of protein (Levitt 1962; Macrae 1970; Powrie 1973; Fennema 1973). This experiment investigated the effect of freezing for 1 d to 12 mo on changes in protein and fibre fractions. In addition, short-term changes (within 1 hour at room temperature) were evaluated to determine if accurate assessment of fresh and ensiled forages is possible. A mechanism for changes that occur during freezing is proposed. 2L1. Ines cutti harv lfloo: HNJS andt follor dean: ofun bags, Hqukfi Specfl fiBezi 61D01 Subs; Ifli7r NaCH) buffer1 (BSCP. 59 2. Experimental 2.1. Treatment of samples Fresh smooth bromegrass (Bromus inermis Leyss.) in vegetative stage of second cutting, and fresh luceme (Medicago sativa) in early bloom of second cutting were harvested from university plots on different days in June 1989. Ensiled luceme (early bloom, second cutting) was collected in June 1989 from a university tower silo after 10 mo storage as silage. Samples were chopped to 3 cm length with a paper cutter, mixed and divided into 30 sub-samples (40 g 12) which were randomly assigned to 10 treatments. Protein solubility fractions were determined for each forage 40 min following harvest or collection from the silo (t=0 h, n=3), 1 h following the first determination (t=1 h, n=3), and following freezing by two methods for several lengths of time (n=24). The samples which were frozen were placed in doubled Whirl-PakTM bags. Half of these bags (n=12) were frozen immediately (t = 0 h) by submersion in liquid nitrogen for 15 min. These and the remaining bags (n=12) were then placed in a p freezer at -25 °C. Several workers processed samples but each was assigned to a , specific task (e. g. cutting, sampling, freezing, analyzing). Samples (n=3) from each freezing treatment were thawed for 15 min i1 at 25° C before analysis after 1 d, 1 mo, , 6 mo or 12 mo of frozen storage. 2.2. Analyses ‘ Sub-samples of fresh or thawed forages (2 g DM basis) were homogenized in 100 ml of ~ pH 7.0 saline—phosphate buffer (4.7 g 1-1 NazHPO4, 2.3 g 1-1 NaH2P04, 4.9 g 1-1 ' NaCl) for 2 minutes, and strained through cheese cloth after 15 min at 25° C. This buffer is similar in pH and osmotic activity to rumen fluid. Buffer soluble crude protein (BSCP) was determined by centrifuging (1500 g, 15 min) 30 m1 of filtrate and analyzing the supernatant for nitrogen. Trichloroacetic acid soluble crude protein (ISCP) we 15 sec, wa nitrogen ] Kjeldahl p emulsion 1 ADCP. T for 6 mo a drying at DM basis fractions 1 1970) wit Z-ethoxyt sulfite (R nimgent analysis 1 NDCP at Cellulose 1982) as C E Grand I carried 0 qualitati' 60 (TSCP) was determined by adding 10 ml of 24% (w/v) TCA to 30 ml filtrate, vortexing 15 sec, waiting 15 min, centrifuging (1500 g, 15 min), and analyzing the supernatant for nitrogen. Nitrogen content of liquid fractions was determined using a modified Kjeldahl procedure (Watkins en! 1987) with the addition of 1 ml 5% Sigma anti-foam emulsion A and a boiling chip to reduce foaming. After sampling for determination of the soluble protein concentration, the remaining sample was dried at 55 °C in a forced air oven for 48 h, ground with a Wiley mill (1 mm screen), and analyzed for DM. ash, NDF, ADF, lignin, CP, NDCP, and ADCP. The first 6 treatments were randomized for these analyses while samples frozen for 6 mo and 1 yr were each analyzed separately. Dry matter was determined after drying at 100 °C in a forced air oven for 16 hours, and all analyses are expressed on a DM basis unless otherwise indicated. Samples were ashed at 500 °C for 5 h. Fibre fractions (NDF, ADF, lignin) were determined sequentially (Goering and Van Soest 1970) with modification of NDF procedure by the substitution of triethylene glycol for 2-ethoxyethanol (18 g kg' 1), and the omission of decahydronaptlralene and sodium sulfite (Robertson and Van Soest 1977). Crude protein was determined as Kjeldahl nitrogen times 6.25 (Watkins M 1987). Residues from non-sequential detergent analysis were collected on filter paper and nitrogen content determined to calculate NDCP and ADCP expressed as a percentage of sample DM. Cellulose and hemicellulose estimates were based on detergent analysis (Van Soest, 1982) as follows: Cellulose = ADF - lignin Hemicellulose = NDF - ADF - NDCP + ADCP. Ground orchardgrass hay was used as a standard for each analysis and all analyses were carried out by the same person. Effect of freezing before drying was also examined qualitatively with light and electron microscopy. Liihmh Data were DM and a1 ash basis, samples. -' (e0hhm were detei Method 0‘ standards of the star 141%“ An additi drying (11 C. Addit compare describet buffer us Models 1 3.1. Cu Ematest bmmegr NDCP t indicatit freezing 61 2.3. Statistical Data were analyzed by the General Linear Models procedure of SAS (1985) on both a DM and an ash basis by forage type using orthogonal contrasts. For calculations on an ash basis, values were divided by percentage ash. Fresh were compared to frozen samples. To evaluate immediate changes after harvesting, the first treatment of fresh (t=0 h) was compared to the second treatment (t=1 h). Duration of freezing effects were determined (1 (1 vs. 1, 6 and 12 mo; 1 mo vs. 6 and 12 mo; and 6 mo vs. 12 mo). Method of freezing (fast vs. slow) was compared within each time period. Dry standards did not vary over time for any analyses except BSCP; therefore mean values of the standards were used as covariates for statistics on BSCP. 2.4. Freeze-drying An additional experiment compared drying at 55 °C without freezing (n=6) to freeze- drying (n=2) for NDCP levels of each forage. Freeze-dried samples were frozen at -25° C. Additionally, bromegrass analyzed for NDCP directly after freeze-drying (n=2) was compared to freeze-dried sample that had been resolubilized in phosphate buffer (as 1 described previously) for 2 h (n=2), and sample resolubilized in reduced phosphate buffer using 3 mg ml‘1 NazS (n=2). The data were analyzed using SAS General Linear Models procedure for unbalanced non-orthogonal contrasts. 3. Results and Discussion 3.1. Crude protein fractions Freezing affected solubility of CP in different solvents (Tables 1-3) with the greatest changes in forage composition observed for protein fractions of smooth bromegrass (Table 1). There was a 41% reduction in BSCP and a 96% increase in NDCP by freezing for l (1 compared to fresh samples. The change in ADCP was slight, indicating the protein which became insoluble in buffer and neutral detergent due to freezing was not made insoluble in acid. There was a subsequent decline in NDCP when freer similar to 1 however, I solubility t in CP and CP percen unfrozen 1 3.2. Ash, Ash perce For exam for l d (T is intuitiv temperan 1'espiratm would be darnage 1 frozen sa lucemec Percenta randomi 1 (Tables for 24 h NDCP; LikGWis attributg 62 when freezing for longer duration. Changes in BSCP and NDCP of fresh luceme were similar to those of bromegrass, but less dramatic (Table 2). In contrast to fresh forages, rowever, there was a slight decline in NDCP due to freezing for ensiled luceme. Buffer solubility of luceme silage declined when stored for 6 mo or longer (Table 3). Changes m CP and TSCP for all forages were small but occasionally significant. The decline in SP percentage for fresh luceme appears to result from a greater respiration of the infrozen material, as CP did not change relative to ash. 3.2. Ash, fibre and lignin Ash percentage of the fresh forages appeared to vary depending on duration of freezing. For example, ash percentage of fresh luceme decreased from 7.2 to 6.9 due to freezing for 1 d (Table 2). Ash is that material which remains after heating at 500° C for 5 h. It .s intuitively unlikely that this material could be lost from the feed as gas at much lower emperatures. Changes in the percentage ash are therefore most likely due to respiratory losses of other feed components. The most likely time for such losses would be during the thawing or 55 °C drying periods. Freezing or thawing may ‘iamage catabolic enzymes and therefore result in decreased respiration during drying of frozen. samples. The increase in ash from 6 to 12 mo of freezing of bromegrass and *uceme contradicts this explanation. For some treatments, the variation in ash )ercentage could be due to randomization error as the last 4 treatments were not 'andomized with the first 6 treatments. Fibre fractions of fresh forages also changed due to freezing for 24 h or longer Tables 1-3). There was an 11% increase in NDF content of bromegrass after freezing 'or 24 h (Table 1 and Fig 2), and 73% of this increase was attributed to an increase in tlDCP; however, estimates of hemicellulose and lignin also increased significantly. hikewise, freezing increased NDF of fresh luceme (Fig 2), and this increase was also .ttributed to increased NDCP, hemicellulose and lignin. The slight increase in NDF due to free: estimate in Sm -25 °C W61 treatments rates of in (contrast 3 coincident significan analyzing 3.3. Visu Frozen sa freeze dri Particles differenc Simply ft in surfac more 310 Howeve 34. Fl‘t Oven (11 Percent; (3.1% 0 that We Freem- Silage ( 63 due to freezing of luceme silage, was attributed to an increase in the hemicellulose estimate in spite of decreased NDCP and lignin. Small differences due to freezing rapidly in liquid nitrogen vs. more slowly at -25 °C were noted for ADCP and TSCP for bromegrass when comparing the 2 treatments within the 24 h time period. Because differences that resulted from different rates of freezing were infrequent, only contrasts from the 24 h time period are shown (contrast 3 in Tables 1—3). Statistical significance of these contrasts may be coincidental given the large number of contrasts made, and there is little practical significance due to the small magnitude of the differences. There was no effect of analyzing immediately after harvest compared to a l h delay at room temperature. 3.3. Visual differences Frozen samples were darker in color and browned compared to samples dried fresh or freeze dried for the fresh forages but did not appear to differ for the ensiled luceme. Particles from these samples appeared thicker using light microscopy, but these differences may have resulted from deposits inside or outside of vascular tissue or simply from a color change. Scanning electron microscopy did not reveal any changes 1 in surface morphology. Bromegrass samples that had been frozen and thawed filtered more slowly after refluxing in neutral detergent than samples freeze dried or not frozen. However, all samples filtered adequately before cooling. 3.4. Freeze-drying Oven dried and freeze-dried treatments were significantly different in NDCP percentage for all forages (P<0.05). Freeze-drying decreased NDCP of bromegrass (3.1% of DM ) compared to oven drying without prior freezing (4.9%). Fresh luceme _ that was freeze—dried had lower NDCP (2.7 %) than oven dried luceme (4.0 %). Freeze—dried luceme silage had slightly lower NDCP (2.6 %) compared to oven dried silage (3.1%). These results are similar to a previous study on freeze-dried and oven dried fora: resolubiliz freeze-drit content of dried we study did protein pr 3.5. Poss Results 01 precipitat Freeze-d1 liquid wa formatior increased carbohyd vacuoles cOIIlplex Sumere 5 increase( may be t or mace iIlsolublt importer DOtentiaj. 64 dried forages (Abdalla c431 1988). In addition, freeze-dried bromegrass that was resolubilized in saline-phosphate buffer before analysis had higher NDCP (5.6%) than freeze-dried bromegrass analyzed directly without resolubilization (3.1%). The NDCP content of bromegrass resolubilized in reduced buffer (3.1%) did not differ from freeze- dried sample analyzed directly (3.1%). Standard errors of the treatment means for this study did not differ significantly and averaged 0.17%. These results suggest that forage protein precipitates in the presence of liquid water after being frozen. 3.5. Possible mechanism Results of these experiments suggest NDCP and other insoluble compounds are precipitated during oven-drying, especially if the samples are previously stored frozen. Freeze-drying appears to inhibit the reaction by preventing transport of reactants via liquid water in frozen and dry samples. High reducing potential seems to inhibit formation of NDCP. Given these characteristics, a possible mechanism for the increased NDCP after freezing and thawing may include the polymerization of proteins, carbohydrates and phenolic compounds. Peroxidase enzymes may be released from vacuoles that are ruptured by water crystals during freezing. These enzymes may complex proteins and carbohydrates into acid-stable and acid-labile complexes (Van Sumere M 1975; Whitrnore 197 8). The acid stable complexes would result in increased ADCP while the acid labile complexes would only increase the NDCP, and may be broken down in the abomasum under acidic stomach conditions. The enzymes or reactants may become less active over time, thus resulting in decreased formation of insoluble material with longer duration of frozen storage. The reaction may be less important to ensiled materials where enzymes have been damaged and reducing potential is higher. 3.6. Role Freezing l affect fibr in forage 1 temperatu degradabi Lignin an NDCP an work has compared Mo 1989] between 1 frozen ant 65 3.6. Relevance of research Freezing has previously been shown to change protein solubility, but was found not to affect fibre fractions of forages (Mme et al. 1975). More recently, however, changes in forage fibre composition have been demonstrated for short-term storage at several temperatures (O'Neil and Allen 1990). Carbohydrate and protein digestibility and degradability may also differ due to changes in chemical composition as observed here. Lignin and ADCP are negatively related to forage digestibility, and greater portions of NDCP and less BSCP may result in less readily degradable protein in forages. Other work has shown faster ruminal degradation of DM and CP from freeze-dried feeds compared to oven-dried feeds for samples submerged in the rumen in dacron bags (V ik- Mo 1989), and in vivo differences in digestion of perennial ryegrass were noted between fresh and frozen samples (Beever 9:31 1974). Fresh forages should not be frozen and thawed before analysis of fibre and nitrogen fractions. Abdalla 1 fresh for. Beever l perennia Fennem: offend: Inc, Nev Natio Wash 4. References Abdalla H 0, Fox D G, Van Soest P J 1988 An evaluation of methods for preserving fresh forage samples before protein fraction determinations. LAW 66 2646-2649. Beever D E, Cammell S B, Wallace A 1974 The digestion of fresh, frozen and dried perennial ryegrass. W 33 73A-74A. Fennema O R 1973 Nature of the freezing process. In. W W ed Fennema O R, Powrie W D, Marth E H. Marcel Dekker Inc, New York, pp 150-239. Goering H K, Van Soest P J 1970 Forage fiber analyses (apparatus, reagents, procedures, and some applications). Agricultural Handbook Number 379. Agricultural Research Service USDA. Jung n G 1989 Forage lignins and their effects on fiber digestibility. Agmnl 81 33-33. Levitt J 1962 A sulfhydryl-disulfide hypothesis of frost injury and resistance in plants. .I W 3 355-391. Lindberg J E 1988 Influence of cutting time and N fertilization on the nutritive value of timotléy8 2. .sgstimates of rumen degradability of nitrogenous compounds. W Res 1 5- MacRae J C 1970 Changes ln chemical corrapositsign of freeze-stored herbage. New ZealandloumalangrieulluraLReseamb 1 45 MacRae J C 1975 Changes in the biochemical composition of herbage upon freezing and thawing. W 84 125-131. Mahadevan S, Erfle J D, Sauer F D 1980 Degradation of soluble and insoluble proteins by Bacteroides amylophilus protease and by rumen microorganisms. .LAnim Sci 50 723-728. Mertens D R 1973 Application of theoretical mathematical models to cell wall digestion and forage intake in ruminants. Ph.D. Dissertation Cornell University, Ithaca, NY. Mertens D R 1987 Predicting intake and digestibility using mathematical models of ruminal function. .LAninLSm'. 64 1548-1558. National Research Council 1985 W National Academy Press, Washington, D. C. Nocek l E. insoluble r Sgi66 166 O'Neil K1 temperatlt Powrie W freeze-pm Fennema Robertsor ! . S . SAS User Van Soes by min. mm Van Soc: of heatin 48 785-7 Van Soc 7594. Van Soe Canine Van Sur Phenolit Sumere Vik~Mc and fro: Watkin sample Weiss ‘ damagt Whitm 241-24 Wood] Organi feedin 67 Nocek J E, Herbein J H, Polan C E 1983 Total amino acid release rates of soluble and insoluble protein fractions and concentrate feedstuffs by Streptomyces griseus. 1D mg Sci 66 1663- 1667. O'Neil K A, Allen, M S 1990 Forage fiber values can be affected by storage time and temperature before drying. JD airy Sci Suppl 7312.8 Powrie W D 1973 Characteristics of food phytosystems and their behavior duringr freeze- -preservationIn: Lw —T m r P rv i n fF n Livin M ed Fennema O R, Powrie W D, Marth E H. Marcel Dekker Inc, New York, pp 352- 386. Robertson J B, Van Soest P J 1977 Dietary fiber estimation in concentrate feedstuffs. J Anim Sci 1Snppl 45 254. SAS User's Guide. 1985 SAS Inst. Inc. Carry, NC. Van Soest P J 1965a Symposium on factors influencing the voluntary intake of herbage by ruminants: Voluntary intake in relation to chemical composition and digestibility ,1 mm; 24 834-843. Van Soest P J 1965b Use of detergents in analysis of fibrous feeds. IH. Study of effects of heating and dryng on yield of fiber and lignin in forages. J Asscc Qffic Anal thm 48 7 85-790. Van Soest P J 1982 Chapter 6: Analytical systems for evaluation of feeds. In: Micrcbia! Ecclcgy cf thc Ruminant, ed Van Soest, P J. O and E Books, Inc. Corvallis, Oregon, pp 75-94. Van Soest P J, Sniffen C J 1984 Nitrogen fractions in NDF and ADF. Distillcrs Fccg anfcrcncc, Cincinnati OH 39 73-82. Van Sumere C F, Albrecht J, Dedonder A, de Poorer H, P6 I 1975 Plant proteins and phenolics. In: The thmistg and Bicchcmistry cf Plant Prctcins, ed Harbome J B, Van Sumere C F. Academic Press Inc, London, pp 210-264. Vik- Mo L 1989 Degradability of forages 1n sacco. 1. Grass crops and silages after oven and freeze drying. Acta Agrjc Scand 39 43- 52. Watkins K L, Veum T L, Krause G F 1987 Total nitrogen determination of various sample types: A comparison of the Hach, Kjeltec, and Kjeldahl methods. 1A sscc Qffic AaLChm 70 410- 412. Weiss W P, Conrad H R, Shockey W L 1986 Digestibility of nitrogen in heat— damaged alfalfa. J Dairy Sci 69 2658-2670. EVhitmore F W 1978 Lignin-protein complex catalyzed by peroxidase. Plant Sci Q11 13 41—245. , Woodman H E, Stewart J 1932 The mechanism of cellulose digestion in the ruminant organism. The action of cellulose-splitting bacteria on the fiber of certain typical feeding stuffs. ,1 Agric Sci 22 527-547. manages» mane. 5..me r mayo" on mama—i o: 3.8 3083 E5 38 35:25 on «885 :83» «3% Am So «L 02:. EB new? menu. EFL? u u E :5 one Baal BEE—en helm». z n z n IE 9% no.» u; N: a.» a: e: we... we... N3 N5 .3 5 new .5 Bu :5 6... :8 Ga Eu :6 Sb 6... .8 2m we do? 5 ..u ..o C 5 i S C C C .8 2m rue wmQum PM 96 ab Pm Pc arm PM u.» wb Pm .3 farm {warm 28E 9— 9.» ch PM uh uh uh uh flu SN .3 Turn rwk >001” _.~ 90 EN Hb Pm Pm fin fin Pm ab .3 Farah warub 22a do.» do.» So one 2b 25 o: 2.: 3a are .3 C in .53 a: 8.» 3... 8e are 8e Ba u; u; 3d .2 _ rub Readies 2.» NS Sb NZ u: N; Sq Sn Ba Ba .3 re 5 8:38." 22 N3 8.» 3e 23 N3 web N3 N3 2... B we we was. we we 3 3 u... u... we we .2 we be 5.3 rub a: 5 So 3 o.» 9e 9e 3 9e 3 .5 .8 we -- £8398 anon am an gases" 825m. cam—.583 8:55 ~88 9.8 25388 on a 3 Base. 3am Qnabuv. axon? CZ 305 5 ~89 9“... .0... 0.3 0.... 00... 0... 0.... 03 m... 0... B. .00 20 . 55.80%... .3 5.0 .w. .3 .3 a... .3 .3 ..3 ..3 .00 3 3 85.08 03 .3 8.0 N... 0.... ..3 0.... 8.0 8... N... 3.. 20 20 0...... q 0 3 3 3 3 3 3 3 3 3 .0. .43 . 8.. . N 3 a... 3 a. 3 3 a... a. 3 ..N 3 -- 3.00080 0.8. o. 0.0 .8280... .0008. 5.8.0.00... 80.8.... ..8... 00.0 0......30 0.. 0 0Q 80:0. 000.0 %Aobmv. 0x00... UK 0.30.... 5 $0... <0. 0.000... N. .no <0. .n... 00.8.“ 8 . 0 .5. 0. 0. 5 Bo 805%. .3 . :8 <0. 0 000 5 Bo $000.5“ 8 0 <0. 5 So 3005.... 8 .88.. 8.00;. <0. 3.02... 50...... N. 0. 0%....008... 000.808 00 0030 0.80... 8.00.30 0.. 0.. 00.. 000.0 QAobmv. 0$30.. 80.0... .0 .550 3.8%... 0330.. 0.02... 0. .8 on .0 $020.. 5a. 30.8.. m0800 08.0... n 3.8%.. x 0.8. 0.08000800000 00.0 30.020 Ow. 5.80.5.0 0.5.0. 8.00.0 Ow. 0.200.... 00.050... 0.50.5.0 0... 5.0.0 00.080... 0.00.0000 n... .2005: 00.05%... 08.0. 050.0 00.0.m0... 00.0. 70 .5me w. Q80. 0. 0800.8 0.. 0800 08.0... 8.0 .....8 080.800 0. .8080 00.80 8 .oo 8-. U7... E .0800. ...0an mmgm. $80.88.“ u .. n... ... .0.. 3.... .N ...0 0.02... 06>... l0 8. 2 .u z .u 2 m 02.. ..3. ..0... ...... ...... ...3 ...... ..3... ..30 ..3.. ...N .00 .... 000 N.... N... N... N3 NN.0 N... NN.. N... N3 N... N. 20 20 0.00.... .03 .0.. .0 .03 .... .00 .00 .0.. .3 ..0 .N3 .3 .3 000.... ...3 ...3 ...3 .30 .33 .3... ...3 ..3 ..... ...0 .00 .. .. 200... ..N 0.. 0.. N0 3 N. 3 3 N.. N.. .... . . >00... ... .0 ... ... ... .3 ... 3 ..0 .0 ..0 3.3 3.3 2.0... 030 000 0... ...0 03 0.... 003 03 003 .03 .NN .3... .. >00... N30 N33 N3 N3. N3.N N33 N30 N30 N3. 030 .N0 0... 20 0.008080. .00 .00 .0... .00 .0.. .0.. ...0 .00 .0.. .0.. ..3 ..0 . 800.000 ...0 ...0 ...0 .... ...N ..... N0.N .... ..3 ...3 ..3 3.3 20 00.... 0.0 03 3.. 3.. 0.0 ..N 0.. 3.0 3... 00 .0. .3... .3 00.. 3 3 3 3 3 ... ... ... ... 0.0 ..N 20 -- N8.00.008 0.8. 0. 0.0 .8080... 800.00. $80500... 00080.0 ..8... 00... 8.00800 0.. 0 08 80.8. 00.0.0 @08me 0x00... UK :00...” C .80.. <0. ..88... 8 We <0. .n... 00...... w. . 0 <0. .. 0. .N ...0 3000.8. .3 . ...0 <0. 0 ....0 .N ...0 0800.8. 3. 0 <0. 5 ...0 3800.8. 0. .88.. 80.0... <0. 0.02... $0.0... NA ... 880.88.. 00080.0 00 80<0 080... 800.800 0.. .... 00.. 000.0 $qu.90.. 0.88.. 80.0..< ... .580 ....880... 0.088.. 0.02.... 0. .3 on ... 0808.. .0.... 80.8.. 80800 08.0... H 0.880.. x 0.8. 080008800000 00.0 00.0.0.0 O... .2880»... 0008. 00.8.0 0... .2008. 00.080... 800.030 Ow. 58.0 00.080... 80080.0 O... .2005... 00.080... ...08. 8.8.0 00.080... 088. DEGM A batch f feeds ust microbia or incorr limiting and free protein ‘ contami pellet tc bacteria finsing NaSOg compa incoqy tWice 1 differe incor; lower degra 71 CHAPTER 3: OBSERVATIONS ON DETERMINATION OF PROTEIN DEGRADATION OF FEEDS BY IN VITRO BATCH FERMENTATION WITH 35SULFUR USED TO LABEL MICROBIAL GROWTH. ABSTRACT A batch fermentation method was developed to measure crude protein degradation of feeds using 358-cysteine to label microbial protein. Previous attempts to use 358 as microbial marker involved inefficient or expensive methods to isolate incorporated 358, or incorrect estimation of N: 358 in microbes. Feeds were incubated (38 °C) in N- limiting media under C02 with rumen microbial inocula. Following fermentation, feed and free bacteria were separated by differential centrifugation. Undegraded feed protein was determined as crude protein of feed pellet corrected for microbial contamination. The Nz358 of microbial pellet was multiplied by 358 content of feed pellet to obtain microbial N. Several methods of rinsing non-incorporated 358 from bacterial and feed pellets were compared with best results obtained by immediately rinsing at least twice in 1% HCl (30 ml) with 1 g/L of each: cysteine.HCl, Na2804, and NaSO3. Several methods of solubilizing microbial pellet to determine 358 were compared with adequate results obtained by solubilizing in water. Feed sulfur was incorporated into microbial protein, thereby increasing Nz358 of bacteria by more than twice that of bacteria grown on NH4. Since feeds differ in S available to microbes, differential centrifugation of each treatment was required for determination of marker incorporation. Using a higher level (3 vs. 6 vs. 11 mg N) of soybean meal resulted in a lower percentage degradation at 20 h. Fraction of feed protein remaining after 20 h degradation was greater (P<.05; n=3) for corn (.40) and meat and bone meal (.32) than for soybe standard publishe treatmen (Keywo protein) Abbrev CP = or DTI‘ = nitroger The rat for the 1989). by fen al., 19: than it severa finely: Physir incub meas feet PTOd' 72 for soybean meal (.12), alfalfa hay (.11) and NH4CO3 (.01). Average within treatment standard deviation was .079. The proposed method is simpler than previously published methods using 358 as microbial marker, however high variation within treatments persists. (Keywords: in vitro protein degradation, 35Sulfur incorporation, rumen microbial protein) Abbreviation Key: CP = crude protein, CTAB = cytyl trimethylammonium bromide, DM = dry matter, DTT = dithiolthreotol, EDTA = ctlry‘ " ' ‘ ‘ aacetic acid, NPN = non-protein nitrogen, TCA = trichloroacetic acid. INTRODUCTION The rate and extent of protein degradation in the rumen is an important consideration for the optimal feeding of high-producing ruminants (National Research Council, 1989). Researchers have developed methods to measure protein degradation of feeds by fermentation with live microbes from the rumen (Henderickx et al., 1972; Raab et al., 1983; Thomsen, 1985; Spencer et al., 1988). In vitro techniques are less expensive than in vivo and in situ methods and they offer the potential to analyze and to compare ‘ several feeds at once. In vitro techniques also allow the researcher to collect and to l analyze both the residue and metabolites of microbial degradation, while more physiological techniques are complicated by fluxes out of the rumen or out of in situ l incubation bags. Methods that involve fermentation with live microbes offer an opportunity to measure protein degradation in an environment similar to that of the rumen. However, these methods require estimation or separation of microbial protein associated with the feedstuff. The extent of protein degradation at a point in time is equal to the sum production of non-protein nitrogen (NPN) and microbial protein. Various procedures have beer digesta. (Walker Miller, 1 sulfide (l pyruvate synthesi protein associat nitrogen entirely associat 0f mien Pr compar fennent detenni An add methoc after a [hill 83} reQuin ‘0 be n micro] 73 have been used to determine the proportion of microbial N associated with feed in . digesta. ‘ Radioactive sulfur (358) often is used to label microbial protein in digesta ‘ (Walker and Nader, 1968; Beever, et al., 1974; Walker and Nader, 1975; Mathers and V Miller, 1980). Microorganisms bring about the enzymatic reduction of sulfate ($04) to , sulfide (HzS) which is then converted to the thiol group of cysteine by reaction with . pyruvate and ammonia or alternatively with serine (Lehninger, 1975). Methionine is synthesized from cysteine, and these amino acids are incorporated into microbial protein. Therefore, incubation of microbes with inorganic sulfur results in radioactivity associated with microbial protein (Emery, et al., 1957). See Figure l. A ratio of nitrogen to radioactive sulfur (N: 358) can be determined for a standard which is entirely of microbial matter. This ratio can then be multiplied by the amount of 358 associated with a mixture of dietary and microbial protein. The product is the estimate of microbial N. Previous methods to label microbial protein have been too labor-intensive to compare a large number of feeds at once. Residual feed protein from microbial . fermentation is digested in nitric or perchloric acid for solubilization of 358 before , determination of radioactivity (Walker and Nader, 1968; Mathers and Miller, 1980). An additional fraction is digested separately for N determination using the Kjeldahl . method. Previously, as much as 20% of the radioactivity has been associated with feed ; after a few hours incubation in sterile media (Walker and Nader, 1968). This indicates 4 that separation of incorporated and non-incorporated 35$ is inadequate. Determination of the ratio of Nz358 also has been labor intensive and has I required acceptance of several assumptions. This ratio is determined on what is hoped to be microbial protein that is not contaminated with feed protein. The standard microbial protein may be obtained by isolating protein from an in vitro rumen incubatior be separat 1975). In the micro Ar microbial feeds W01 accompli that have incorpor. al., 1972 compare '1 method 1 methods Separatit determir degmda sources not enti 80 that : BXperie 74 incubation that used only NPN (Henderickx et al., 1972). Alternatively, microbes may be separated from dietary protein by differential centrifugation (Walker and Nader, 1975). In either case, one assumes that the protein isolated has the same N23SS ratio as the microbial protein associated with the feed sample. An early goal of the present study was to develop a system to measure microbial growth using an abundance 35S-cysteine, so that contribution of sulfur from feeds would be negligible compared to that included in the system. If this goal were accomplished, the Nz3SS ratio might be constant across all treatments and standards that have only soluble protein. (NPN). Cysteine was desired as a marker because direct incorporation of feed amino acids may be favored over inorganic sulfur (Henderickx et al., 1972), thus lowering the incorporation of 358 of microbes incubated with feeds compared to those incubated with blanks. The objective of this study was to simplify an in vitro microbial fermentation method to measure protein degradation of feeds. This required simplification of methods to solubilize microbial protein for detection of radioactivity, a more thorough separation of incorporated and non-incorporated radioactivity, and simple but accurate determination of Nz358. The method was evaluated by determining fractional degradation of protein supplied at different levels, comparison of several protein sources and the variation associated with those measurements. These objectives were not entirely met, but the results of several experiments are presented in this dissertation so that anyone wishing to continue this work will have the benefit of the these experiences. cl 75 SO3— sulfate 802' sulfite S(s) <—> H28 elemental sulfide sulfur pyruvate + NH3 NH3—CH-COOH —> NH3-CH—COOH | CH2-SH CH2—CH2—S-CH3 cysteine methionine Microbial Protein Figure 1. Sulfur Metabolism in the Rumen. Final Metht Freeze-drier mm screen I drying in a‘ a modified In V media is bl bicarbonat digest of 0 make the 1 concentra‘ microbial carbohyd addition . microbia protein, 5 at .2 m1] dairy co microbi Ilnmedi Cyseir dithiolt E tWice t “final- 76 METHODS Final Method for Microbial Fermentation Freeze-dried forages and dry concentrates were ground through a 5 mm and then a 1 mm screen of a Wiley mill. Dry matter (DM) was determined by weighing hot after drying in a forced air oven for 8 h at 100 °C. Crude protein (CP) was determined using a modified Kjeldahl method (Hach et al., 1985). . In vitro media was adapted from that of Goering and Van Soest (1970). This media is buffered with NaHCO3 and includes macro- and micro-minerals. Ammonium bicarbonate was replaced with an equi-molar concentration of NaHCO3, and trypsic digest of casein (trypticase, Sigma Chemical Co., Saint Louis, MO) was excluded to make the media low in N. Magnesium Sulfate was replaced with an equi-molar concentration of MgC12.6H20 to prevent a decrease in specific activity of 358 in microbial protein due to dilution with cold sulfur from added sulfate. Soluble carbohydrates were added in the quantities shown (Chapter 4, Table 1). Vitamin addition was adapted from Russell et al. (1983). In order to decrease affects on microbial growth, and hence protein degradation, due to amino acid profile of added protein, isoacids (valerate, isovalerate, isobutyrate and 2-methyl-butyrate) were added at .2 ml/L of media. This rich media was used to simulate that of the high-producing dairy cow fed a mixed diet, and to remove constraints to protein degradation (i.e. low microbial growth) that are unrelated to protein structure of the feed to be tested. Immediately before use, media was purged with C02 until the pH stabilized at 6.8. Cysteine.HCl (.05 g, except where indicated otherwise) and 353-1abe1ed L-cysteine in dithiolthreotol (DTT, Aldrich Chem. Co., Milwaukee, WI) was mixed into the media. Rumen fluid was collected 2 to 4 h after feeding from a Holstein dry cow fed twice daily a diet based on corn silage and alfalfa hay. Rumen fluid was blended for 1 min and strained through 4 layers of cheese cloth followed by glass wool. Rumen fluid was incub method (it Feet tubes. Tu 25 m1 of 1 incubated At the en with Ant temporar In l min. Th Trichlon supernat consider (20.000 consider supema combin. of the f. in the ri Sample time in labeled incorp. I Suspt‘n 77 was incubated in continuous culture at a dilution rate of .07/h for 24 h according to the method described elsewhere, except media was as described previously. Feed samples containing 4 to 6 mg N were placed in 50 ml plastic centrifuge tubes. Tubes were sealed with black rubber stoppers connected to a source of CO2, and 25 ml of labeled media and 8 ml of rumen inocula were added. Samples were incubated at 38 °C while in equilibrium with 1 atm C02 for varying amounts of time. At the end of the incubation, degradation was arrested by addition of 2 ml of 6 N HCl with Antifoam Emulsion A (Sigma Chemical Co., Saint Louis, MO) and sample storage temporarily at 4 °C. In the final procedure, digested feed samples were centrifuged at 400 g for 30 min. The supernatant was decanted and centrifuged again at 20,000 g for 30 min. Trichloroacetic acid (TCA, 10 ml per 30 ml sample) was added to this second supernatant which was centrifuged a third time (1500 g, 20 min). The first pellet was considered residual feed protein, bound bacteria and protozoa. The second pellet (20,000 g) was considered to be composed only of free bacteria. The third pellet was considered to be composed of soluble feed and released microbial protein. The final supernatant was discarded as degraded protein. The first and third pellets were combined and rinsed with 1% (vol/vol) HCl solution (30 ml) containing 1 g/L of each of the following: cysteineHCl, NaSOg, and Na2804. Sample was allowed to soak in the rinse solution for l h after which time TCA (10 ml, 240 g/L) was added and samples were centrifuged at 1500 g. The pellet was rinsed and precipitated a second time in the same manner. Three replicates of feed were also incubated for 3 h in labeled media with no inocula to determine amount of activity that precipitates without incorporation in microbial CP. Rinsed pellets were suspended in H20 (10 m1), and an aliquot (.2 ml) of this was suspended in scintillation cocktail (BioFluorTM, New England Nuclear, Dupont Co., Boston, l Searle A1 standard samples : sample tl remaindt pellet). Develop Several protein 1 were co was add should 1 Cellulo. h witho precipit several methoc Resuln in plac Variant disinte quencl detem sol/lit: 78 Boston, MA). Radioactivity was determined in a scintillation counter (Isocap 300, Searle Analytic Inc.). Quench was determined from channel ratios determined on standard curves with differing amounts of soybean meal added. Radioactivity in samples incubated without microbial inocula was subtracted from the respective feed sample that had been inoculated and incubated. Crude protein was determined on the remainder of the sample as Kjeldahl N times 6.25. Undegraded feed CP was estimated as: Feed pellet CP - (353 in feed pellet (x) CP in bacterial pellet/ 35$ in bacterial pellet). Development and Validation Several experiments were conducted to determine the best method to separate microbial protein from the original label. Most of these experiments (unless indicated otherwise) were conducted as those involving microbial fermentation except no microbial inocula was added. Incubation with labeled media without the addition of microbial inocula should result in no precipitation of 358 because there should be no microbial growth. Cellulose (160 mg), corn starch (200 mg) and soybean meal (100 mg) were incubated 4 h without addition of rumen inocula. Following incubation, samples were directly precipitated with TCA (rather than centrifuging out the bacterial pellet). Because l several experiments were conducted with variations to the methods used (as the final method was not yet developed), these methods are described in more detail in the Results and Discussion section of this paper. Generally, replicate values are presented in place of means because the sample size was small for these experiments, and the variance was not uniform across treatments. Radioactivity is presented as disintegrations per minute unless quench was not determined (assuming uniformity of quenching across treatments) in which case counts per min are given. Quench was ' determined by channel ratios adjusted to a standard curve when quench was added from soybean meal, or by internal standards when indicated. Sev detection cysteine experimt digest ot TCA pn peptide Smgd Results amount bacteri: ccntrifr previor solven one ha meat 2 run. I fenne Sepa non. addi‘ Dtec 79 Several experiments were conducted to compare methods to solubilize 358 for detection in the supernatant. These methods included comparing recovery of 35S cysteine from solutions as well as detection of activity in labeled microbial samples. Radio-labeled microbial protein was isolated from feed protein in two different experiments. In the first experiment, microbes were grown in the presence of trypsic digest of casein (Sigma Chemical Co.), or NH4Cl. The ratio 35$:N was determined for TCA precipitate of this degradation after incubation for 18 h with different amounts of peptide or ammonia N added initially (0, 1, 3, or 5 mg), and with incubations with 0 or 5 mg of peptide or ammonia N for different periods of time (0, 3, 9, 18, or 48 h). Results were analyzed for consistency of 353 to ratio over time and with different amounts of substrate. In the second experiment, 358 to N ratio was determined on the bacterial pellet from degradation of feeds. The free bacteria were isolated by centrifugation at 20,000 g after removal of feed that precipitates at 400 g, as described previously. Feeds used and their respective percentage CP content on 21 DM basis were: solvent extracted soybean (Glycine max) meal -- 48.0, alfalfa (Medicago sativa) hay at one half bloom —- 19.1, dried and ground corn (Zea mays L.) grain -- 10.8, and dried meat and bone meal —- 55.2. Three replicates of each feed were incubated in the same run. Protein degradation was determined by the final method for microbial fermentation described previously for the 5 feed samples. RESULTS AND DISCUSSION Separation of Incorporated from Non-Incorporated 35$ 1 ' n f r u in n ri r rin in Rumen bacteria may synthesize non-protein sulfur compounds such as sulfide (82'). This experiment was to see if addition of Na2S is necessary to remove sulfide to prevent its association with precipitable protein. Trypticase (61 mg), Starch (200 mg) and cellulose (160 mg) were digested i 1i)6 (1me of N328}. incubatic cocktail t than whr would bt would b other su‘: 80 digested 10 h with rumen inocula and media containing cysteine (1 g/L) and 35S (8.2 x 106 dpm) to produce labeled microbial protein and sulfide. Treatments were addition of Na2S.9H20 (40 mg), NazS.9H20 and HCl (1 ml 6 N) or no addition at end of incubation and before rinsing once in water. Pellet was suspended in scintillation cocktail (10 ml). Total radioactivity (cpm) in pellet was greater for NazS treatments than when none was added as indicated in Table 1. One would expect that H28 gas would be removed from the system by addition of acid and that more radioactivity would be remOved by addition of Na2S. However, the reducing agent may have made other sulfur compounds more likely to bind when rinsed in distilled water. A second experiment was conducted at the same time to determine the effect of additions prior to rinsing when no fermentation has occurred (only cysteine in the system). Cellulose (160 mg), starch (200 mg) and soybean meal (100 mg) were incubated 4 h without addition of rumen inocula. Treatments were addition of NazS (40 mg), ‘2-mercaptoethanol (5 m1), DTT (50 mg), or no addition prior to TCA precipitation and rinsing three times with water. Precipitation of 35S (dpm) was much greater for all treatments than for the previous experiment, but there were no differences among treatments. The greater precipitation of 35S in this experiment compared to the previous one may be explained by greater binding to soybean meal than trypticase. Total radioactivity in pellet is given in Table 2. Pmcipitatign of cysteine with protein, This experiment was conducted to compare different methods of precipitating protein for their effects on precipitation of 35S from cysteine. Trichloroacetic acid, tungstic acid, perchloric acid and copper sulfate precipitation were compared, each at two concentrations of precipitating agent (5 or 10 ml of reagent to 30 ml buffer). Each treatment was centrifuged (1500 x g; 15 min) or filtered (Whatman #1), and was carried out with cellulose (160 mg) and starch (200 mg) and no protein or with these carbohydrates and soybean meal (100 mg). Mel Cor 81 Media (30ml) with 358-cysteine (1.6 x 106 dpm) was added to centrifuge tubes, and precipitating reagent added immediately. There was no added cold cysteine. Concentrations of reagents were 24% TCA, 20% perchloric acid, .05 M copper sulfate and 1 N sodium hydroxide, and 10% sodium tungstate and 1.07 N sulfuric acid. Table 1. Precipitation of 358 (cpm) with trichloroacetic acid after incubation of Trypticase with microbial inocula for 10 h with addition of sulfide or sulfide and acid before rinsing the pellet in water. Treatment Rep 1 Rep 2 Rep 3 Na28 25,960 27,362 17,077 . Na2S + HCl 48,492 34,368 26,181 No addition 12,977 14,491 16,032 Table 2. Precipitation of 358 (cpm) with trichloroacetic acid after incubation of soybean meal with sterile media for 4 h with addition of reducing agents before rinsing the pellet in water three times. Treatment Rep 1 Rep 2 Rep 3 NaZS 95,143 86,896 101,271 2-mercaptoethanol 94,468 100,169 D'IT 84,245 108,337 103,942 No addition 109,500 34,215 89,564 Tungstic : pellet wa in Table i 4 filter p 4 filter p presence did not : precipit determi to deter of mic] cystein and 1 and 1t (10 m'. 15 mi Shake TCA suSpe rinsi ante Was dlfft 82 Tungstic acid and copper sulfate precipitations used 2.5 or 5 ml of each reagent. Total pellet was suspended in scintillation cocktail (10 ml), and radioactivity (cpm) is given in Table 3. Tungstic acid preparations would not filter through Whatman # 1 so Whatman # 4 filter paper was used. Copper sulfate preparations filtered poorly through Whatman # 4 filter paper as well. Cysteine associated with the precipitate in all cases. The presence of soybean meal appeared to increase precipitation of cysteine. These results did not suggest much improvement would be attained by changing from the TCA precipitation procedure. Rinsing the pellet of LES—cysteine precipitate. This experiment was to determine the best method to rinse non-protein 35S-cysteine from TCA precipitate, and to determine if an adequate amount of 35S could be removed for accurate determination of microbial sulfur. Rinse solutions were: A. water, B. cysteine HCl (1.3 mg/ml), or C. cysteine.HCl and Na2$ (1.3 mg/ml). Samples were washed once in solutions A and B and 1, 2 or 3 times in solution c. In vitro media (30 ml) with 353-cysteine (4.1 x 106) and 10 mg or 0.5 mg unlabeled cysteine was added to soybean meal (100 mg). TCA (10 ml) was added after a 15 min delay. Samples were shaken and centrifuged (1500 g; 15 min), and supernatant discarded. Rinse solution (30 ml) was added and samples shaken. After 15 min delay (30 min delay for one treatment with rinse C one time), TCA was again added and centrifugation and decanting repeated. Total pellet was suspended in scintillation cocktail (10 ml), and radioactivity (cpm) is shown (Table 4). Rinsing in cysteine solution was more effective at removing 35S-cysteine than rinsing in water. There was no apparent improvement by adding Na28. Rinsing twice appeared better than once and three times was better than twice. However, this effect was only apparent at higher level of initial unlabeled cysteine. Considering the 20 fold difference in initial cold cysteine concentration with little difference in precipitation of Table 3. protein, ‘ Ira-1| 1313M Water. 83 Table 3. Precipitation of 35S-cysteine (cpm) with different methods to precipitate protein, with and in the absence of soybean meal.1 Reagent SBM Conc. Method rep 1 rep 2 Trichloroacetic acid 0 10 ml centrifuge 23,062 32,706 filter 8,625 9,403 5 m1 centrifuge 28,218 —-— filter 11,874 10,552 100 mg 10 m1 centrifuge 70,629 49,630 filter 35,041 72,682 5 ml centrifuge 23,818 94,512 filter 31,396 30,904 Tungstic acid 10 ml centrifuge 38,897 45,840 filter 36,095 25,806 5 m1 centrifuge 32,120 38,343 filter 41,017 24,625 100 mg 10 m1 centrifuge 50,076 44,462 filter (#54) 50,366 45,648 5 m1 centrifuge 103,501 86,995 filter (#54) 70,695 Perchloric acid 10 ml centrifuge 43,364 38,550 filter 39,716 28,184 5 ml centrifuge 43,951 47,077 filter 80,730 66,210 100 mg 10 ml centrifuge 30,984 32,531 filter 99,133 82,333 5 ml centrifuge 51,304 49,821 filter Copper sulfate 0 10 ml centrifuge 73,897 filter 23,376 10,864 5 m1 centrifuge 58,059 44,936 filter 21,641 15,269 100 mg 10 ml centrifuge 28,204 26,635 filter (#54) 23,496 5 ml centrifuge 23,31 l 40,529 1SBM = soybean meal, conc. = concentration of reagent (5 or 10 m1 used with 30 ml water. Tablt (cpm Initi.‘ lOn 84 Table 4. Method of rinsing the trichloroacetic acid pellet on removal of 35S—cysteine (cpm) from a feed mixture (starch, cellulose and soybean meal). Initial Cys Rinse Times Delay r9) 1 rep 2 10 mg Water 1 15 min 8038 6407 Cys 1 15 min 5843 2991 Cys+Na2S 1 15 min 6925 4479 Cys+Na2S 2 15 min 3813 3623 Cys+Na2S 3 15 min 3117 3172 Cys+Na2$ 1 30 min 2605 6965 0.5 mg Water 1 15 min 7975 5895 Cys l 15 min 3191 Cys+Na2S 1 15 min 3661 4079 Cys+Na2S 2 15 min 2731 7875 Cys+Na2$ 3 15 min 6261 8468 Cys+Na2S l 30 min 8798 35S t cystei theon There effect and c conta mg) i incub TCA 33ml) deter show orgat solut incul from 85 358, either a constant amount of 35S precipitates or a constant percentage of total cysteine. Decreasing the unlabeled cysteine in the ill vitro media would therefore, theoretically increase the ratio of labeled protein to interference from non-protein 358. Therefore, if low levels of unlabeled cysteine are used in the media, 35S should be effective in labeling microbial S. Another experiment compared rinsing 1 time with a solution containing Na28 and cysteine, compared to water or no rinsing. Samples were incubated in media containing 35S-cysteine (5.0 x 105 dpm) for 8 h with starch (200 mg), cellulose (160 mg) and trypticase (62 mg) or soybean meal (100 mg). Samples with trypticase were incubated with live rumen organisms while soybean meal was not inoculated. After TCA precipitation and rinsing in 30 ml solution and TCA precipitation repeated, samples were digested in H2804 with perchloric acid (Hach et al., 1985), quench was determined on external standards with H2304, and total pellet radioactivity (dpm) is shown in Table 5. These results confirmed the previous experiment for incubation without live organisms. Rinsing in H20 reduced dpm compared to not rinsing, and rinsing in a solution containing cysteine and Na2S reduced dpm compared to rinsing in H2O for incubations with soybean meal and trypticase. Percentage activity precipitated ranged from 1.5 to 5.3. An experiment compared rinsing l, 2 or 3 times in solution (30 ml) containing 1 g/L of each: Na2SO4, NaSO3, Na2S and cysteine-HG]. Soybean meal, starch and cellulose were incubated for 3 h in media containing cysteine (.2 g/L) and 3SS-cysteine (3.0 x 105 dpm), and without rumen bacteria. After TCA precipitation, addition of . rinse and repeated precipitation in TCA, the pellet was digested in H2SO4 with peroxide. Total radioactivity precipitated is given Table 6. Table 5. for 8 h v ’ Treattnt Trypticz No 1 H2( Cys Soybea No 86 Table 5. Precipitation of 35S (dpm) from media incubated with rumen microorganisms for 8 h with Trypticase, or without organisms with soybean meal. Treatment Rep 1 RELZ W Trypticase No rinse 24,340 18,085 26,383 H20 rinse 12,298 18,383 18,426 Cys-Na2S 16,936 14,255 13,702 Soybean meal No rinse 16,128 20,723 26,553 H20 rinse 15,191 19,021 17,787 Cys-Na2S 7,277 7,745 9,787 Fl‘otal activity in each tube was 5.0 x lOme. Table 6. Effect of number of times of rinsing a feed sample on removal of 35S-cysteine from TCA precipitable material.1 Rinse Rep 1 Rep 2 Rep 3 1 time 7,106 12,681 7,489 2 times 1 4,809 4,979 5,617 3 times 4,383 4,383 4,043 1Total activity in each tube was 3.0 x 105 dpm. Table 7. Effect of soaking feed sample during rinsing on precipitation 35S-cysteine.1 Soak Rep 1 Rep 2 R933 0 h 1477 3191 407 8 2 h 1950 2246 2128 20 h 2 187 2009 2305 1Rinse solution was a .1 M sodium phosphate buffer with 1 g/L of each cysteine.HCl, Na28, Na2$04, NaSO3. Total activity per tube was 3.6 x 103 dpm. sul in 87 Table 8. 355-cysteine precipitated with different feed samples incubated (38 °C) for 3 h without microbial inoculal Treatment rep 1 rep 2 rpp 3 Soybean meal 4,213 4,766 5,191 Bromegrass 13,361 13,787 15,702 Starch 2,085 1,872 2,979 Cellulose 4,340 4,043 5,021 flbean—acid rinsed 1,191 1,957 2,383 1Each sample was rinsed 3 times in pH 7 phosphate buffer (.1 M), with 1% KBr and 1 g/l each of the following: cysteine-HCl, Na2SO4, NaSO3 and Na2S, or in pH 1.5 dilute sulfuric acid. Total activity placed in each tube was 5.2 x 106. Additional rinses reduced the precipitation of 355, but the difference between 2 and 3 rinses was not great. Another consistent trend is becoming apparent. The reduction in precipitated 358 due to rinsing procedures is accompanied by a reduction in variation among samples within a treatment. An experiment was conducted to determine the effect of allowing the sample to soak in rinse before adding TCA. Each sample was rinsed twice and the pellet was allowed to soak in the rinse for 0, 2 or 20 h before addition of TCA for the second rinse. The pellet was made radioactive as described in the previous experiment except 3.6 x 105 dpm 358 was added. The rinse solution was the same as that above except it was made in a .1 M phosphate buffer (pH 7). There appeared to be a reduction in 358 precipitation due to soaking for 2 h as compared to 0 h. However, there was no effect of soaking for 20 h compared to 2 11. Another experiment compared substrates for their ability to precipitate 35S. Soybean meal (100 mg), smooth bromegrass hay (300 mg), starch (200 mg), or cellulos microbi in the in pH 7 ph cysteine in dilute Bach trt 113A fo cysteine precipit through to reduc better u (pH 1.5 contain Sllpema cenm‘fu and the measun Supema CYSteint That is the Star 88 cellulose (160 mg) were incubated with 358-cysteine (5.2 x 106 dpm) without microbial inocula for 3 h at 38 °C. Amounts of substrate were used that would be used in the in vitro procedure for protein degradation (4 to 6 mg N). Samples were rinsed in pH 7 phosphate buffer (.1 M), with 1% KBr and 1 g/l each of the following: cysteineHCl, Na2804, NaSO3 and Na28. One u'eatment of soybean meal was rinsed in dilute sulfuric acid (pH 1.5) instead of phosphate buffer and with the other reagents. Each treatment was rinsed three times as before with a 16 h delay before addition of TCA for the final rinse. Total radioactivity (dpm) is shown in Table 8. These results show that different substrates result in different amounts of 35S- cysteine precipitated. This should emphasize the importance of minimizing the precipitation of non-microbial 358 because simply subtracting a constant for all feeds throughout the degradation may result in systematic bias. Rinsing in acid also appeared to reduce the precipitation compared to neutral buffer. A rapid method to measure 358 precipitation was used in several experiments to better understand the association of 358-cysteine with the pellet. Dilute sulfuric acid (pH 1.5. 30 ml) containing 358—cysteine (6.8 x 103 dpm) was added to centrifuge tubes containing starch (200 mg). The samples were centrifuged (15 min; 1500 x g) and supernatant discarded. Rinse solution (20 ml) was added to the samples followed by centrifugation. This was repeated twice. The final suspension (before centrifugation) and the final supernatant were sampled (.2 ml) for radioactivity. Precipitated 35S was measured as the difference in radioactivity between the final suspension and the final supernatant. The first use of this technique showed that starch pellet that already had 358- cysteine associated with it did not result in additional precipitation of 35S-cysteine. That is it appeared that the starch was saturated with radioactive cysteine. However, the starch Table Treatr pH 1 pH 7 M Table trimetl starch. Treatm EDTA DTT CTAB Control may hat OCCUITBI shown 1 have 8hr c)lStine i 89 Table 9. Effect of rinse solution pH on precipitation of 35S—cysteine (cpm) after rinsing twice. Treatment rep 1 rep 2 pH 1 3 10 pH 7 295 250 pH 10.6 80 25 Table 10. Effect of adding -11 J ‘ " ' ‘ ‘ " acid, dithiolthreotol, or cytyl trimethylammonium bromide to rinse solution on 35S-cysteine precipitation (cpm) with starch. Treatment Rep 1 Rep 2 Rep 3 EDTA 164 88 157 DTT 247 239 162 CTAB 205 178 245 Control 218 242 258 may have been reduced after the first precipitation so that no appreciable oxidation occurred the second time. The effect of pH was investigated for its affect on the association with starch as shown in dpm below. The association was inhibited by acid compared to pH 7. Others have shown than the oxidation of cysteine to cystine occurs fastest at neutral pH, and cystine is insoluble. Low pH rinse solution resulted in the lowest precipitation of 35S. The pellet from pH 7 treatment was then rinsed 3 times in dilute sulfuric acid (pH 1.5) to see if the acid rinse could reverse the binding. There was no decrease in binding due to rinsing in dilute acid. Rinsing an additional time in pH 0.5 sulfuric acid also did not reduce the association of 35$ with starch. Rinsing again by soaking for 2 h in pH 0.5 solution also did not remove 35S from the pellet. However, nearly all 358 was s (repli fOI'IIII ethyl (CTr stare The ' and I 2 ml were the i scin tend no t solu beft botl witl eitl. cen tin: 90 was solubilized by rinsing in 1% HBr solution brought to pH 7 by phosphate buffer (replicated twice). The Br' ions may have been able to substitute for bound cysteine formed in this particular case. Further exploration of use of KBr may be desired, but HBr vapors are hazardous and so use of both acid rinse and HBr may not be desired. An experiment was conducted to determine the effect of adding etlr,‘ " ' ‘ ‘ ancetic acid (EDTA), DTT, or cytyl trimethylammonium bromide (CTAB) to rinsing the pellet of 353-cysteine. Samples included .17 g cellulose, .2 g starch and .062 g soybean meal and were incubated for 4 h without microorganisms. The Control rinse contained 10 ml/L HCl and 1 g/L of each: cysteine-HCl, Na2804, l and NaSO3. Treatments were addition of: EDTA at 3 mM final concentration, D'I'I‘ at l 2 mM final concentration, and CTAB at 8 mM final concentration. The treatments were for the first rinse only, and a second rinse with control rinse solution was used for the second rinse. Samples were dissolved in 40% H2SO4 and diluted before scintillation counting. Total cpm are shown in Table 10 for each of three replicates. The results show a high variation in precipitation of 353. There was a tendency for the EDTA treatments to be lower in precipitable 35$ than the control with no effect of the other treatments. The EDTA did not dissolve in the acidic rinse solution. An experiment was conducted to determine whether there is precipitable 353 before the incubation, and if so whether it could be precipitated before the experiment. A year old source of 35S-cysteine was used for the experiment. The media and rinse both contained 1% HCl to prevent oxidation of cysteine. Samples were rinsed twice with the solution used previously. The treatments included centrifugation of the media either in the presence or absence of starch (.2 g) and cellulose (.16 g). After centrifugation the supernatant was added to starch, cellulose and soybean meal and rinsed twice as previously. An additional treatment was prepared the same way but 01 de of di; ad Ti Wt dir in' inc 91 contained 5 times more cold cysteine in the media. The control treatment was not centrifuged to remove precipitable material before adding to feed sample and rinsed. Each treatment was replicated 3 times. There were no significant differences among treatments and 0.8 % of the total radioactivity precipitated ill each case. The experiment was repeated the next day using the same isotope. This time the precipitate ranged from 0.45 % of the total for samples subjected to centrifugation before rinsing, compared to 1.5% for samples not subjected to centrifugation before rinsing. Addition of 5 times more cold cysteine also reduced the precipitation of 35$ to 0.18% of the total. The results suggest that some precipitable 358 can be removed by centrifugation before incubation. If some ferrnentations result in greater precipitation of non-microbial 35S than others, then bias is introduced by this precipitation. An experiment was conducted to determine the effect on precipitation of 35S due to the reduction procedure and length of incubation. Media was degassed by heating to boiling and cooling under C02. In retrospect, this was not a very efficient way to degas media compared to using a bubble disperser. Media was reduced with cysteine-HCI (1 g/L ) and 35S-cysteine (2.3 x 106 dpm/tube) was added. Feed (.1 g soybean meal, .2 g starch and .16 g cellulose) was added to centrifuge tubes and 30 ml media added to the feed and incubated at 38 °C. Timing of reduction, addition of 35S, addition to the feed and precipitation in TCA were varied as indicated in Table 11. Each sample was rinsed twice in 1% HCl with 1 g/l each of the following: cysteine-HCI, Na2SO4, NaSO3, and radioactivity determined directly on a sample (.2 ml) of a suspension of the final pellet. At the same time, some samples were incubated with 1% HCl in place of in vitro media, these results are shown in Table 12. The most important effect to be determined from this experiment is that incubation without organisms results in precipitation of 358 over time. The longer the rn at 92 incubation, the more likely is the precipitation. It could not be prevented by reducing conditions or even acid. Most researchers have used 35SO4 to label microbes. This compound may be less likely to adhere to feeds. However, it is readily converted to SZ‘ which, like cysteine, is quickly bound to feed samples (Walker and Nader, 1968). Solubilization of Incorporated 358 Tissue solubilizer and Hg] digestion. Commercial tissue solubilizer (BTS-450, Beckman Co., Fullerton, CA) and 6 N HCl were used separately to solubilize the pellet from TCA precipitate. Due to the high amount of carbohydrate in the samples both approaches left a black viscous residue that resulted in high (>70%) quench. Sample size was reduced from entire pellet to .2 ml resolubilized pellet (resolubilized in 10 m1 H20). This procedure resulted in an apparently adequate digestion in BTS-450 with low quench, as will be discussed later. Hg :1 digestion and chromatography with Dowex, The purpose of these experiments was to determine if 35S-protein could be determined by HCl digestion followed by separation using chromatography. Orchard grass hay (.25 g) was digested for 24 h with 6 N HCl with 2—mercaptoethanol (15 ml) at 110 °C. After digestion 35S- cysteine was added to digest. 35S—cysteine in 1 N HCl (n=2) or 6 N HCl digest (n=2) was put on Dowex anion exchange columns (.2 ml). The columns were rinsed with .01 N HCl and cysteine eluted with .5 N NH4OH. Only about 20% of the 353 was recovered from the columns when cysteine was put on the columns directly. Only about 2% of the 35S from the digest was recovered from the columns. Moreover, recovery increased only 20% for fractions of NH4OH over fractions of HCl. Tab] in pl WISE Oh Oh 121 121 93 Table 11. Effects of timing on precipitation of 35S—cysteine (dpm) with feed samples in pH 6.8 media. Media Feed Precip- reduced 353 added added itated Rep 1 Rep 2 Rep 3 +incubated 0 h 0 h 12 h 15 h 13,950 15,550 14,400 0 h 12 h 12 h 15 h 20,700 19,100 14,300 12 h 12 h 12 h 15 h 24,350 14,600 19,600 12 h 12 h 12 h 63 h 39,450 25,700 - 27,600 0 h 12 h 0 h 15 h 7,850 9,050 4,000 0 h 12 h 10 h 12 h 4,050 5,050 3,600 1 0 h 12 h 10 h 15 h 9,400 7,100 6,350 l 0 h 12 h 10 h 36 h 24,600 20,500 22,950 0 h 12 h 10 h 60 h 17,350 32,950 4,300 1Each sample was rinsed 3 times in 1% HCl with 1 g/l each of the following: cysteine-HCl, Na2SO4, NaSO3 and Na2S. Total activity placed in each tube was 2.3 x 106 Table 12. Effects of timing on precipitation of 35S-cysteine (dpm) with feed samples in 1% HCl. Media Feed Precip- reduced 353 added added itated Rep 1 Rep 2 Rep 3 +incubated 0 h 12 h 10 h 12 h 31,750 8,250 9,600 0 h 12 h 10 h 15 h 9,550 30,000 12,550 0 h 12 h 10 h 36 h 21,700 18,050 20,950 0 h 12 h 10 h 60 h 13,250 22,250 18,950 1Each sample was rinsed 3 times in 1% BC] with 1 g/l each of the following: cysteine-HCI, Na2SO4, NaSO3 and Na2S. Total activity placed in each tube was 2.3 x 106 5.6‘ the dirt Six in 1 COT firs we sec W3 0X ha lal tht 94 A second attempt was made to elute 358-cysteine from anion exchange columns (n=2). This time a higher concentration of NH40H (50%) was used. The first HCl fraction from 358.cysteine in 1 N HCl contained 48% of the original 353. The first Nmon fraction (3 ml) contained 15% of the original 358, with an additional 6.2% recovered in other fractions. When the 358-cysteine was added with acid digest (n=2), 5.6% eluted with acid wash and 5.2% eluted in the first NH40H fraction and 3.5% in the remaining two fractions. The attempt to elute cysteine was repeated by putting 35S-cysteine in l N HCl directly on the columns (n=2). This time .1 N HCl was used as rinse instead of .01 N. Sixty-two percent of the 358 activity was recovered that was recovered was recovered in the elute. The first HCl rinse fraction contained 22.3% and the first NH40H fraction contained 26.9%. The experiment was repeated and similar results were obtained. The first NH40H fraction was added to .1 N HCl and put on new columns. Two peaks were again found except that this time they were in the second fraction of HCl and second fraction of NH40H. The BC] peak was 3.8 times higher than the NH40H peak. It is possible that the cysteine was oxidized to cystine or degraded once the pH was neutralized with NH4OH. Cystine is insoluble which could explain why low recoveries were found. Addition of reducing agents and an anaerobic environment, or oxidation to cysteic acid, may have improved the recovery. The cysteine may not have had time or surface area enough to bind to the column thus explaining the recovery in the HCl fraction. The attempt to use ncr digestion and chromatography to solubilize 358 in the pellet was discontinued for several reasons. The procedure was found to be more laborious than necessary. Much of the 358 in cysteine appeared to bind irreversibly to the column. The 358 in cysteine did not elute in a single fraction. There was 95 considerable interference from other compounds in the HCl digest with elution on the column. W. The first experiment was conducted to determine if digestion in concentrated sulfuric acid with H202 could be used for determination of 35S in TCA precipitate. This would allow digestion for radioactivity to be simultaneous to digestion for ninogen determination by Nessler's reaction. However, N and S differ in that N is non-gaseous in acid even when boiling at high temperatures. 0n the other hand S can be expected to evaporate with sulfuric acid. This experiment measured the amount of 358 present after digestion in H2804. The amount of H2804 remaining was determined by titration with 8 N NaOH. An equal amount of 358 was added to each of 5 flasks. Orchard grass (.25 g) was added to flasks 1 to 3, and H2804 (4 ml) was added to each additional flask. The first 5 flasks were digested for varying amounts of time to digest sample and allow some evaporation. The flasks were titrated with TKN solution as indicator, then were filled to 100 ml with H20. Samples (.2 ml) for 358 activity were taken. Samples that were digested resulted in lower acidity and lower cpm than the undigested sample. This suggests that 35S disappeared with the reduction in acid. However, the sample that was not digested had a higher ratio of cpm/acidity, suggesting that more cpm were lost from samples that were digested than acidity was lost. Therefore, on the basis of one sample, this may not be an appropriate way to digest samples to determine 358 activity. On the other hand, the cpm per ml NaOH required to neutralize acid was constant for the other samples. Perhaps, activity lost in these samples was from the sulfide contamination rather than cysteine. Further work with more replication is required to reevaluate these preliminary results. Addition of 358- cysteine to 1 ml 20% H2804 resulted in 94% efficiency of scintillation counting {I'll-Ir}! l? 96 compared to the same radioactivity in 20 m1 aqueous environment. This indicates that the quenching from acid is not a serious problem to overcome. A different approach to estimating H2804 remaining after digestion was investigated. The concentrated acid was sampled after digestion for radioactivity. The flask was then filled to 100 ml with H20 and sampled again for 353 and N. The acid remaining after digestion was calculated as follows: Conc. H2804 = (100 ml x dpm/ml of diluted acid) / (dpm/ml of cone. acid) Theamount of acid could then be used to calculate the percentage reduction in N from sampling the concentrated acid for radioactivity. The sampling of concentrated acid was highly variable with a conventional repeater pipette, but was improved substantially by using a positive displacement pipette. Sampling of concentrated H2804 was still variable because of the viscosity of the material. Alternatively, the amount of acid remaining after digestion was determined by weight. This approach assumes that only acid remains after the digestion process. An experiment was conducted to see if 358 from free cysteine was lost proportionally to H2804 during digestion. Four treatments were compared: 1) no digestion of H2804, 2) boiled in H2804 for 10 min., 3) boiled in H2804 for 10 min and 10 m1 H202 added, and 4) boiled with in H2804 with 20 ml H20 and l Kemwipe with H202 added for complete digestion. The replicate weights of the residues and cpm in 0.2 ml of digest diluted to 100 ml with H20 are presented. The initial weight of the acid varied from 7.27 to 7.39 with all but one sample between 7.36 and 7.39. Weight of the residual acid appeared to increase due to digestion for treatment 2 (Table 14). The digestion decreased weight of residual for treatment 4, presumably to evaporation of sulfuric acid. Digestion (treatment 2-4) decreased the recovery of cpm. There was a greater decrease in cpm than in weight. This could be due to an incorrect measurement of acid or to a greater loss of 358 than of total S. In either case, the 97 digestion in sulfuric acid resulted in a loss of 35S that could not be easily accounted for. Use of weight to determine evaporation of H2804 would be further complicated if feed samples contain ash that increases weight of digested samples. With concentrate feeds, sample size was small relative to the amount of acid, which would decrease this error. Additional experiments demonstrated that digestion in cold concentrated H2804 resulted in high quenching (> 50%) even if diluted to 20% acid for scintillation counting. The acid accounted for 4% of the quenching, with the remainder due to the coloration of acid digested samples. Quenching could be reduced by the addition of H202 for decolorizing, but addition of .3 ml 50% H202 to .2 ml 20% H2804 resulted in 20 % quenching from H202. The quenching (as determined by internal standards) from color and H202 was reduced by incubation in the oven for 30 min at 55 °C. and was reduced more by incubation at 100 °C. However, this treatment resulted in an unexplained loss of 358 activity. The final method of solubilizing the microbial pellet was. to suspend sample in 6 m1 H20 and add 4 ml concentrated H2804. The solutions are rapidly mixed to prevent coloration of the sample. This mixture is diluted 50% before scintillation counting to reduce acid quenching and facilitate dissolution in cocktail. All microbial protein was dissolved in this solution as was purified starch and cellulose. However, some feed protein remained undissolved. 98 Table 13. Recovery of 358 (cpm) in samples digested in H2SO4 with peroxide and recovery per ml NaOH required to neutralize the H2SO4 in the digest. NaOH (m1) cpm cRm/NaOH 16.0 1735 108 17.2 1828 106 14.5 1671 105 18.6 2025 109 17.2 1904 110 21.0* 2771 131 *Last sample was not digested. Table 14. Digestion of 35S-cysteine in sulfuric acid and peroxide on recovery of radioactivity (cpm). Treatment Residual mass cpm cpm/mass Not digested 7 .27 g 4803 661 7.36 4899 666 Boiled in H2SO4 7.73 g 4000 517 7.76 4256 548 H2S04 + H202 7.33 g 4409 602 7.53 4605 612 Digested Kemwipe 5.77 g 2529 438 6.71 3032 452 99 WW. An experiment was conducted to determine if the efficiency of measuring 353 radioactivity in microbial protein directly without digestion. Treatments were: 1) radio-labeled microbes suspended in water and suspended in scintillation cocktail, 2) radio-labeled microbes suspended in 20% (vol/vol) H2804, and 3) radio-labeled microbes solubilized in BTS-450. There were 3 replicates per treatment. Microbes were by incubating with rumen inocula at 38 °C for 17 h with trypticase (3 mg/ml) and carbohydrates in media. The microbes were precipitated with TCA and rinsed twice using the final method determined previously. Quench was determined by internal standard. Though treatment with 20% H2804 appeared to decrease recovery of 358 from microbes compared to solubilizing in water or BTS-450, treatments appeared equal for effect of quenching determined on internal standards. These results suggest that use of BTS-450 is not necessary for detection of 358 activity. However, perhaps none of the treatments solubilized all of the activity in microbial samples. These treatments should be compared to samples treated by complete acid digestion to ensure complete recovery. Secondly, a problem with use of 358 is the high variability within treatments. A larger number of samples per treatment is required to evaluate effects of these solubilizing treatments on variability. Ratio or N to 35s in Labeled Microbes Effects of length of incubation on microbial growth are demonstrated in Table 16. Microbial inocula was incubated for 24 h in continuous culture before addition to centrifuge tubes with media and carbohydrates. This inocula was already radio-labeled when added to the tubes. Microbial growth appeared to cease within 3 h of incubation for both ammonia and trypticase. Though 5 mg of N was added to each tube, microbial pellet only contained 3.5 mg N. Perhaps some microbial N was solubilized in TCA at 100 the end of the incubation. The 35S to N ratio increased over time for NH4 but remained fairly constant for trypticase. The preincubation media had trypticase in it. This media thus supplied additional sulfur to growing microbes which diluted the incorporation of 358. Additional growth on NH4 resulted in an increase in N to 358 ratio because less sulfur was available in the media. This ratio continued to increase up to 48 h even though additional microbial growth plateaued at 3 h. It is possible that more 35S precipitated with the pellet without incorporation, but this did not occur for the samples incubated in trypticase, so this is an unlikely explanation. Alternatively, microbial protein may have been degraded by microbes and resynthesized resulting in more microbial 35S incorporated over time. In any system in which the amount of S in the incubation is different from that of the preincubation, there may be a change in 35S:N over the length of the incubation, thus requiring that this ratio be determined for each time point to which it is to be applied. Table 15. Solubilization of 35S from radio-labeled microbes using different solvents. Treatment dpm* EfficielgLl Solubilized in water 74835 0.867 Solubilized in 20% H2SO4 69285 0.856 Solubilized in BTS-450 76298 0.853 SBM (n=3) 1576 0.019 *Significant effect determined by ANOVA, P < .05. 1Fraction of internal standard measured in spiked samples. T: ni a"D Table 16. Effect of incubation time with microbial inocula and nitrogen source on nitrogen and 355 activity insoluble in trichloroacetic acid. Time (h) mg N dpm dpm/N NH4 (5 mg N) u u a... 0 1.6 955 617 3 3.0 1915 639 9 3.4 2776 816 18 3.5 3281 954 48 3.2 3583 l 134 SEM (n=3) .1 119 65 Trypticase (5 mg N) ** 0 2.4 995 423 3 4.4 1564 356 9 3.9 1384 353 18 3.9 1627 414 48 3.3 1519 462 SBM (n=3) .5 164 9 Blank HI IHI 0 1.6 826 509 3 1.9 1080 57 1 9 1.9 1 121 594 18 1.7 1263 727 48 1.7 1337 768 SEM (n=3) .1 41 27 P, N sources .0001 .0001 .0001 *Differences among treatments within a row and N sources, P < .05. ** P < .01. Table 17. Effect of amount and source of nitrogen on nitrogen and 35S activity insoluble in trichloroacetic acid after incubation with microbial inocula for 18 h. Added N (m g) %N dpm dpm/N NH4 u u an 0 1.7 1263 727 l 2.2 1341 628 3 2.8 2460 865 5 3.4 3281 954 SEM (n=3) .1 205 95 Trypticase sul- lul- u 0 1.7 1263 727 l 2.3 1479 634 3 3.2 1557 490 5 3.9 1627 414 SEM (n=3) .1 46 16 P, N source .0001 .0001 .001 ”Differences among treatments within a row and N sources, P < .01. 103 The effects of altering the amount of nitrogen incubated with microbes for 18 h are shown in Table 17. The greater the amount of NH4 supplied to the media, the greater the ratio of 35S to N. This effect is due to the increased incorporation of 35S from the media with a higher specific activity than that used in the preincubation as described for the time effects. Incubation with greater amounts of trypticase resulted in lower 35SIN ratio due to the dilution of 35S by S in the trypticase. Again the ratio of 35S to N in the microbes depended on the contribution of S from the nitrogen source. Undegraded CP in different feeds after incubation with ruminal microbes for 20 h as described for the final method to determine protein degradation of feeds (Methods section) is shown in Table 18. Corn grain and meat and bone meal were less degraded than alfalfa hay and soybean meal. These trends are expected (National Research Council, 1989). The ammonia was shown to be completely degraded as is expected. This sample served as a control to evaluate the ability to measure protein degradation and microbial growth. The undegraded protein is presented as the fraction of the initial 5 mg N from feeds incubated with rumen microbes. The amount of undegraded feed CP was much less than the amount of bound microbial protein estimated from 35$. Often when concentrates are evaluated for protein degradation, there is no correction for microbial contamination of the microbial protein associated with feed protein. This correction was shown to be very important to estimation of protein degradation. Data in Table 18 also shows that the level of soybean meal incubated with enzyme affects the fraction degraded. This may be an artifact of the method, because a negative value was obtained at the lower level. Possible differences due to sample size would indicate that first order kinetics do not apply. Further research is needed to determine if it is possible to eliminate this variation due to sample size and to determine a fractional degradation. 104 Table 18. Estimation of microbial nitrogen and undegraded feed protein after incubation for 20 h with rumen inocula. Flee Microbial Pellet Undgr. Bound Feed (fraction) m.o. N N (mg) dpm/N Corn grain _ .40a 2.7c .33c 5489": MBM 323.!) 2.60 .336 475b,c Alfalfa hay .1 lead 3.31).c .51b 444c NH4C03 .01d.c 3.11).c .54b 951a Soybean meal 3 mg N -.0t5e 4.213 .320 604b 5 mg N .1204! 2.7‘3 .53b 537b,c 11 mg N .231),c 6.13 .88a 497bac SBM (n=3) .047 .42 .043 43 arbchifferent superscripts within a column indicate differences between means, P < .05. lUndgr. = Undegraded feed protein (fraction remaining of 5 mg N used at start), Bound m.o. N = microbial nitrogen associated with feed pellets, dpm/N = disintegrations per minute per mg N of microbial pellet (20,000 g), MBM = dried meat and bone meal. 105 Estimation of ”SN varied for different feeds due to the incorporation of feed sulfur into microbial protein. This result means that this ratio must be determined for each. feed analyzed, because this ratio is important to the calculation of microbial protein associated with undegraded feed. Since the amount of feed sulfur available for incorporation into microbes will change over time and with sample size, 3SSIN must be determined for every ueatrnent. Not only does this add to the labor intensity of the method, but the variation associated with determination of ”SM is also transferred to the final measurements. The goal of incorporating 35S into microbial protein in the presence of large amounts of cysteine was not attained. There is always some precipitation of radioactivity with feed protein even when no incorporation has occurred. This precipitation was very high relative to the specific activity of 353 in microbial protein when large amounts of cysteine were used in the incubation. It was possible to effectively label microbial protein, but low amounts of sulfur were required in the media, so that a higher percentage of the 358 was incorporated into microbial protein, and a lower percentage of non-incorporated 35S was precipitated with feed and microbes. Using media low in sulfur resulted in a change in 35S:N for different treatments due to the incorporation of feed sulfur in place of 35S into microbial protein. This work exposes many of the limitations to using 353 to label microbial growth for the determination of protein degradation of feeds. The final method presented here was simplified from previous efforts, but it is still too complicated for analysis of a large number of feeds at once. The variation, even within a run as shown in this paper, was too high to detect any but huge differences between feeds. This variation was due to the fact that calculation of protein degradation was dependent on six different measurements all of which had variation associated with them. Each sample required estimation of: l) 353 activity in feed incubated with no microbial inocula, 2) 353 in 106 microbial pellet, 3) 35S in feed pellet, 4) N in feed pellet, 5) N in microbial pellet, and 6) initial N of feed sample. Some of this variation might be reduced in the future by lowering the precipitation of 35S with feeds and by developing more consistent methods to solubilize the residual feed for determination of 353. Use of sulfate instead of cysteine may result in less precipitation of non-incorporated 358, but sulfate must be converted to cysteine before it can be incorporated into microbial protein. The microbes may preferentially use feed sulfur over sulfate, thus decreasing the incorporation of 35S04. A method is still needed that uses microbial fermentation to measure protein degradation of a large number of feeds, and with the ability to analyze the end-products. The available methods all have many of the limitations shown in this paper. The work described in this paper is limited by the lack of replication, and methods used to rinse and solubilize pellet may have been inadequate for some experiments. Further development of the techniques described here will require the iterative approach used so far. Development of better means to solubilize microbial protein for detection of radioactivity requires a better means to remove non-microbial 35S from samples. Better means to determine 35S2N require better rinsing and solubilizing of microbial protein. The description of the results described in this paper is intended to make it easier to continue development of these methods, and to avoid the pitfalls of the present work. ...l N b.) P E" O\ >‘ 9° 107 REFERENCES . Beever, D. E., D. G. Harrison, D. J. Thomson, S. B. Cammell, and D. F. Osbourn. 1974. A method for the estimation of dietary and microbial protein in duodenal digesta of ruminants. Br. J. Nutr. 32: 99. . Emery, R. S., C. K. Smith, and C. F. Huffman. 1957. Utilization of inorganic sulfate by lumen microorganisms. I. Incorporation of inorganic sulfate into amino acids. Appl. Microbiol. 5: 360. . Goering, H. K. and P. J. Van Soest. 1970. Forage Fiber Analyses (Apparatus, Reagents, Procedures, and Some Applications). Agric. Handbook N0. 379. ARS- USDA, Washington, DC. Hach, C. H., S. V. Brayton, and A. B. Kopelove. 1985. A powerful Kjeldahl nitrogen method using peroxymonosulfuric acid. J. Agric. Food Chem. 33: 1117. Henderickx, H. K., D. I. Demeyer, C. J. Van Nevel. 1972. Problems in estimating microbial protein synthesis in the rumen. Page 57 in Tracer Studies on Non-Protein Nitrogen for Ruminants. Inter. Atomic Energy Agency. Vienna. . Lehninger, A. L. 1975. Biochemistry. The Molecular Basis of Cell Structure and Function. 2nd ed. Worth Publishers, Inc. New York, NY. Mathers, J. C. and E. L. Miller. 1980. A simple procedure using 35S incorporation for the measurement of microbial and undegraded food protein in ruminant digesta. Br. J. Nutr. 43: 503. National Research Council. 1989. Nutrient Requirements of Dairy Cattle. 6th rev. ed. Natl. Acad. Sci., Washington, DC. 108 9. Raab, L, B. Cafantaris, T., and K. H. Menke. 1983. Rumen protein degradation and biosynthesis 1. A new method for determination of protein degradation in rumen fluid in vitro. Br. J. Nutr. 50: 569. 10. Russell, J. B., C. J. Sniffen, and P. J. Van Soest. 1983. Effect of carbohydrate limitation on degradation and utilization of casein by mixed rumen bacteria J. Dairy Sci. 66: 763. 11. Spencer, D., T. J. V. Higgins, M. Freer, H. Dove, and J. B. Coombe. 1988. Monitoring the fate of dietary proteins in rumen fluid using gel electrophoresis. Br. J. Nutr. 60: 241. 12. Thomsen, K. V. 1985. The specific nitrogen requirements of rumen microorganisms. Acta Agric. Scand. Suppl. 25: 125. 13. Walker, D. J. and C. J. Nader. 1968. Method for measuring microbial growth in rumen contents. Appl. Microbiol. 16: 1124. 14. Walker, D. J. and C. J. Nader. 1975. Measurement in vivo of rumen microbial protein synthesis. Aust. J. Agric. Res. 26: 689. 109 CHAPTER 4: EN RICHMENT 0F PROTEOLYTIC ACTIVITY IN PREPARATIONS FROM THE RUMEN FOR IN VIIRO STUDIES Swords Three experiments compared treatment of rumen contents on recovery of proteolytic activity per ml rumen fluid (PA) and per mg nitrogen (N). Proteolytic activity was determined by degradation of azocascin for 2 h, and N was determined by Kjeldahl method. The first experiment compared effects of chilling, blending, rinsing, fractionating by centrifugation at 400 x g and 20,000 x g, and their interactions. The second experiment compared rumen fluid from a f'lstulated cow to the same rumen fluid that was incubated in continuous culture (dilution rate = .07Ih) in media with high sugar content. The third experiment compared effects of extracting centrifugation pellets of rumen fluid with buffered media, media with detergent, and acetone. For all experiments, most activity was found in the centrifugation pellets. Rinsing of rumen solids with cold buffer increased PA and PAIN. Chilling tended to increase PA slightly, but had no effect on PAIN. Blending rumen contents did not affect PA, but decreased PAIN. Incubation in continuous culture did not affect PAIN. Mixing rumen fluid precipitate with buffered media solubilized 18% of PA. Treatment with detergent or acetone solubilized 51 and 48 % of PA respectively. Activity per mg N was 3.9 times greater in supernatant than in original sample after treatment with detergent or acetone. Proteolytic activity was enriched most by treatment with detergent or acetone, and additional recovery and enrichment were provided by rinsing of rumen solids. Protein degradation in vitro: Proteolytic activity 110 Introduction Various researchers have developed methods to measure protein degradation of feeds by incubation in vitro with enzymes from rumen organisms. One approach has been to incubate feeds with rumen fluid in the presence of microbial growth inhibitors (Broderick, 1987; Neutze er al. 1993). Another approach has been to extract enzymes from ruminal contents (Mahadevan er al. 1987, Kohn and Allen, 1993). These methods are limited by interference of non-enzymatic proteins in rumen fluid preparation. Contaminating proteins may compete with feed protein for enzyme sites thereby reducing protein degradation of feeds . In addition, contaminants may interfere with the ability to measure degraded or undegraded protein of the feed. Since the contaminant may degrade at a slower rate when in the presence of feed protein due to competition for enzyme, subtraction of values from blank (no feed) samples may not be adequate. Ideally, these methods require that enzyme preparations from the lumen are representative of whole rumen contents, yet do not contain additional proteins. Previously, proteolytic activity in rumen fluid preparations has been increased by chilling rumen contents before straining through cheese cloth, or by rinsing rumen solids in cold buffer, and blending of rumen contents has had inconsistent effects (Craig et al., 1984; Furchtenicht and Broderick, 1987). These treatments may increase proteolytic activity (PA), and still have a negative impact on proteolytic activity per unit protein or nitrogen (N). Sequential centrifugation at 500 x g and 20,000 x g divides rumen fluid into a) soluble contents, b) free bacteria, and c) protozoa, feed and bound bacteria (Kopecny and Wallace, 1982). Understanding the enzyme partitioning in these fractions may elucidate possibilities for enrichment of PA. Sacchrolytic rumen bacteria are primarily responsible for protein degradation (I-Iungate, 1966). These organisms have high reproduction rates (Lin et al., 1985), so incubation in continuous culture at high dilution lll rate and with high availability of saccharides and starch may be another method to increase PA or PAIN of rumen contents. Extraction of enzymes by solubilization in water (Wallace and Kopecny, 1983), and by treatment with detergent (Kopecny and Wallace, 1982) or acetone (Mahadevan et al., 1987) has been successful at solubilizing proteolytic activity, but these methods have not been compared to each other within a single study. Interaction effects of many of these means of enrichment need elucidation to determine which combinations are optimal for recovery and concentration of PA. Azocasein is used to measure proteolytic activity of rumen contents (Mahadevan et al. 1979; Wallace and Kopecny, 1983). Enzymatic degradation of azocasein results in the solubilization of azo dye in trichloroacetic acid (TCA), but this could result from either cleavage of dye from protein or from actual protein degradation. The azo dye may itself be degraded from the activity of rumen contents thus causing an underplediction of activity (Mahadevan et al., 1979). Usually absorption of light at 440 nm wavelength is used to detect azo dye in alkaline solution (Charney and Tomarelli, 1947), but interference from absorption of rumen fluid may complicate this measurement. Rumen fluid has been centrifuged to remove interfering pigments before analysis of azo casein degradation (Mahadevan et al., 1979), but if some proteolytic activity remains soluble, the total activity may be underestimated. The objectives of the present study were to: 1) validate and improve measurement of PA in rumen contents using azocascin, and 2) determine effects of chilling, blending, rinsing, fractionation by centrifugation, incubation in vitro, and extraction of rumen contents on proteolytic activity in vitro. MATERIALS AND METHODS Validation and improvement of azocascin method Color and nitrogen liberation from azocascin during degradation were compared to ensure that solubilization of dye is due to protein hydrolysis (which releases N) and not 112 simply cleavage of dye. Azocasein was solubilized in sodium phosphate buffer (.1 M, pH 7) and nitrogen and optical density at 440 nm (OD) were measured for standards not subjected to TCA precipitation. Centrifuge tubes containing 2 ml phosphate buffer, 16 mg azocascin and 1 m1 rumen inocula were incubated at 38°C for 0, 1, 2, 4, or 8 h. At the end of the incubation, TCA (1 ml) was added to precipitate undigested protein. After centrifugation, OD. and N were measured on 1 ml aliquots. Samples which did not contain azocasein (blanks) were also incubated and azocascin added immediately before addition of TCA. These blanks were used to measure N and absorbance not due to azocasein solubilization, and they were subtracted from the digested azocascin samples. The optimal wavelength to measure light absorption in solutions with azocascin was determined. Strained rumen fluid was centrifuged at 20,000 x g for 30 min. The pellet was resolubilized in an equal volume of water as the initial volume of rumen fluid. Resolubilized rumen fluid pellet, the rumen fluid supernatant, and azocascin solution (.2 mg azocascin I ml water) were each combined with KOH at a final concentration of IN. The absorbance was then determined at wavelengths ranging from 260 to 800 nm at intervals of 20 nm. Effect of adding microbial growth inhibitor to azocascin degradation was determined. Azocasein degradation was determined as described in the final method except media was not reduced and sodium hydrazine (final concentration 2 mM) was added to the inhibited treatment. The inhibitor was chosen because it has previously been shown to inhibit microbial growth at the concentration used without affecting enzymatic activity of proteases (Broderick, 1987). Each treatment was replicated three times with separate batches of rumen fluid. Two experiments compared timing between steps on absorbance of degraded azocascin samples. In the first experiment, KOH was added immediately after 113 centrifugation of final preparation, or after soaking in TCA overnight. Each treatment was replicated three times. In a second experiment, six samples were read for absorbance immediately or after soaking in KOH solution overnight. A precipitate appeared to form in the bottom of tubes soaked in KOH, so the optical density of these samples was measured both after remixing or centrifugation. Azocasein degradation: final method In vitro media used in these experiments except where counter indicated was modified from Goering and Van Soest (1970) by replacing NH4HCO3 with an equimolar concentration of NaHC03. The media was diffused with CO2 gas before use until the pH stabilized at 6.8, and was reduced with .6 g/L of each cysteine-HCI and Na28. Azocasein (8 mg/ml) was solubilized in the in vitro media. Ruminal enzyme source (.5 m1) and media with azocascin (1.5 ml) were added to 15 m1 centrifuge tubes. Tubes were incubated under C02 at 38°C for 2 h. The degradation was then arrested and undigested protein precipitated with addition of ml TCA (240 gfkg water). Blank samples were determined by incubating enzyme preparations without azocascin solution, then adding azocasein immediately before addition of TCA. After addition of TCA, samples were centrifuged (1500 x g, 20 min). An aliquot of supernatant (1 ml) was transferred to another centrifuge tube and 2.5 N KOH (1 ml) was added. Optical density (O.D.) was determined at 480 nm wavelength with a spectrophotometer (Beckman DU 460, 1 cm path length). Optical density of the blank from each sample was subtracted from that of the digested azocascin for that sample. This value was compared to a standard curve derived from dilute azocascin solution in l N KOH. Each step in the procedure was completed without delay to avoid loss of color over time due to TCA or N aOH solutions. 114 Chilling, blending and rinsing Rumen contents were collected from a fistulated Holstein dry cow fed twice daily a ration based on maize (Zea mays L.) silage and luceme (Medicago sativa) hay. Rumen contents were collected two to four hours after feeding and divided into 4 parts. Two parts were chilled for 4 h and two were processed immediately. One part from each initial treatment (chilled vs. not chilled) was blended and one was not. A114 parts were strained through 4 layers of cheese cloth followed by glass wool. Cold (4°C), reduced in vitro media as described for use in azocasein degradation was poured over the filtrand, and the rinse collected and kept separate from the strained fluid that was initially removed. Each treatment was thus divided into two additional fractions (strained fluid and the rinse of contents). These eight samples were centrifuged sequentially at 400 x g for 15 min and 20,000 x g for 30 min and each fraction (supernatant, pellet l and pellet 2) analyzed for N and 2-H azocascin degradation. The experiment was replicated 6 times using different rumen fluid samples from the same cow on different days. Results were analyzed in a split-plot design blocked by rumen fluid sample (run). Chilling and blending treatments were nested within rumen fluid sample. Rumen fluid and the rinse of contents were nested within chilling and blending treatments. Centrifugation fractions (400 g pellet, 20,000 g pellet, supernatant) were nested within rinse and fluid. The mean of PAIN determined in fractions is not proportional to PAIN determined on the whole of a treatment. Therefore, three separate ANOVA's were computed. First, total PAIN for each replicate (n = 6) of chilling and blending treatments was determined by finding the sum of PA in the six fractions (2 rinse x 3 centrifugation fractions), and dividing that value by the sum of N in the six fractions. Chilling, blending and their interaction effects were determined after blocking by run using these total PAIN values. In a second ANOVA, PAIN was 115 determined for each replicate used in comparison of rumen fluid to rinse of solids. Replicate values were calculated as the sum of PA from the three centrifugation fractions divided by the sum of N. Comparison of rumen fluid to rinse and the interactions of this contrast with chilling and blending were determined using the calculated replicates, after blocking for run, chilling, blending and chilling x blending interaction effects on the PAIN calculated the same way. In a third AN OVA, effects of centrifugation fractions and all their interactions with chilling, blending and rinsing were determined on the PAIN determined on the individual rumen fluid fractions. Effects of chilling, blending, rinsing and their interactions and effect of run on PAIN determined on fractions were removed from error. Interactive effects were elucidated by computing the previously described contrasts for each rinsing and centrifugation fraction rather than using the sum of fractions. For this and all subsequent experiments, statistics were performed using JMP (SAS Institute Inc. 1991). Incubation in vitro Cheese cloth filtrate of rumen contents that was blended but not chilled (500 ml) as described for the previous experiment was used. In vitro media as described for azocascin degradation was reduced with .6 g/l Na2S.(H20)9. Carbohydrate (Table l) and tryptic digest of casein (1.5 g/L) were added to select for proteolytic organisms based on recommendations of Pichard (1977). Concentrations of carbohydrate chosen to select for proteolytic organisms. The incubation was carried out in a one liter flask with side arm, under diffusion of C02 and in a 38° C water bath. Media with carbohydrates was added at constant rate of .07/h using a peristaltic pump to perfuse media through glass tubing to the bottom of the flask. Effluent was collected in a separate flask. The contents of the fermentation flask were continually stirred with a magnetic bar. Incubation in continuous culture was continued for 24 h. 116 Both the incubate and the initial rumen fluid were subjected to sequential centrifugation (400 x g and 20,000 x g, 30 min), and the proteolytic activity on azocascin and N content of each fraction was determined. The experiment was repeated on six different days (run). Comparison of initial rumen fluid with infusate was made after blocking by run for each centrifugation fraction and for the total of the three fractions. Extraction in bufl'er, detergent or acetone Different solvents were compared for extraction of enzyme from rumen contents. Rumen fluid was collected and prepared as described for the previous experiment. Strained rumen fluid (50 ml per sample) was cooled to 4 °C and then centrifuged at 20,000 x g for 30 min. These centrifugation pellets were resolubilized in 25 m1 reduced media (previously described), reduced media with detergent (deoxycholate) or acetone. Each sample with solvent was blended for 30 sec in a tissue homogenizer (Sorvall Omni-mixer). Samples treated with acetone were centrifuged again (20,000 x g, 30 min) and acetone soluble fraction was discarded. This pellet was blended with 25 ml reduced media. After resolubilization, all treatments were centrifuged (20,000 x g, 30 min). RESULTS AND DISCUSSION Validation and improvement of azocascin method The ratio of azo dye to N solubilized was constant over time and the same as for the sample not treated with TCA. This indicates that azo dye liberation is the same as N liberation and represents true protein degradation, not merely the cleaving of the azo dye from the protein. If azo dye was degraded by ruminal enzyme as has been recorded for long-term degradation (Mahadevan, 1979), the ratio of dye to N solubilized would decrease over time or differ from the undigested sample. The solubilization of azo dye 117 was linear (constant mg/ml/h) for up to 8 h. Since the results are positive, azocascin degradation was used in several other experiments. The absorbance of light due to rumen fluid pellet and supernatant, and to azocascin is shown in figure 1. The wavelength chosen to measure azocasein absorption for the remaining experiments was 480 nm because, though it was not the wavelength of peak light absorption for azocascin (450 nm), it resulted in less interference from light absorption of rumen fluid than the lower wavelength. There was no effect of adding hydrazine on azocascin degradation (P > .1). The mean degradation was 2.28 mg azocascin hydrolyzed/h/ml rumen fluid and the standard error of the mean was .24. Therefore hydrazine was not added in subsequent experiments. Soaking in TCA overnight resulted in lower (P < .01) OD. than not soaking (.44 vs. 1.1 mg/h). Soaking in KOH overnight did not affect CD. but centrifugation at this step did result in lower 0D. (1.2 vs 1.4 mg/h, P < .01). Because it was determined that solubilized dye faded if left in TCA supernatant addition of KOH was completed without delay and samples were mixed by gentle shaking before analysis (azocascin fragments that are soluble in TCA but precipitable in KOH were considered degraded. Chilling, blending and rinsing Chilling of rumen contents before straining tended to increase PA recovered slightly, but blending had no effect (Tables 2 to 3). Rinsing of rumen contents increased total recovery of PA by 42% (Tables 2 to 3). Chilling increased PA in rumen fluid pellets, but did not affect PA in rinse (Figure 2). This effect explains most of the interaction with chilling and rinsing, and shows that rinsing rumen solids in cold buffer is nearly as good as chilling before straining at recovering enzyme activity. The rinse of rumen solids had higher PAIN than fluid (Tables 2 to 3), indicating that rinsing not only helps in recovery of PA but that it can concentrate that activity. Chilling had no overall effect 118 on PAIN (Table 2), but chilling increased PAIN in the rinse of solids with no impact on the fluid (Fig. 3). Blending decreased PAIN (Tables 2 to 3) and therefore is not recommended as a treatment to concentrate proteolytic activity. There was also a significant chilling by blending interaction on PAIN in the rinse fraction (Table 2). Proteolytic activity differed for centrifugation fractions with most activity recovered in pellets (Tables 2 to 3). Sequential cenuifugation separates rumen fluid proteins into three fractions. The supematant contains free soluble protein. The 20, 000 x g pellet (pellet 1) contains free bacterial protein and the 400 x g pellet (pellet 2) contains protozoa, insoluble feed protein and bound bacteria. Most proteolytic activity was associated with the feed/protozoa and bacterial fractions and little enzyme appears to have been soluble. These results confirm earlier work that showed most proteolytic activity was insoluble (Kopecny and Wallace, 1982). Chilling increased PA in pellets but not supernatant (Tables 4 to 5). Therefore, it did not appear to be such a harsh treatment as to dislodge the enzymes from the bacteria. Much of the activity in the rinse fraction was precipitated at 400 x g, while little was located in the bacterial fraction of the rinse (Tables 4 to 5). This indicates that rinsing may not actually dislodge free bacteria, but rather dislodge protozoa or feed with bound bacteria. Proteolytic activity per unit N was greater in the 400 x g pellet and supernatant of the rinse than for the strained fluid (Tables 4 to 5), thus confirming this hypothesis. Most of this increase was found in the rinse from chilled samples (Figure 3). The present work, as well as previous studies, demonstrates that rinsing of rumen solids is effective at increasing both PA and PAIN. However, none of the effects on concentration of proteolytic enzyme relative to crude protein would be adequate to ensure excess enzyme for extended periods of incubation. Rinse of rumen solids would be important to increase the percentage of activity recovered, and potentially this recovery could impact how representative the recovered enzyme would be. Based on 119 these results, it was clear that both centrifugation pellets would be important to recovery of enzyme, yet since the supernatant represented a small portion of the total activity it could be removed. Incubation in vitro Results of incubation of rumen fluid in continuous culture are shown in Table 6. Preincubation resulted in a build up of brown sludge in the bottom of the flask, demonstrating that solids were not washed out of the flask. This fraction was precipitated into the 500 x g pellet which contained a surprisingly high level of PA. However, PAIN was lower in preincubated contents than for the original for this pellet. The 20,000 x g pellet had lower PA in preincubate than rumen fluid. This would most likely be explained by a dilution of microorganisms in continuous culture owing to the high dilution rate relative to microbial growth. The PAIN was not different for the free bacterial fraction indicating that the free bacteria of preincubate did not have greater capacity to degrade protein than those bacteria in the original fluid. The PA in the supernatant was also lower in preincubate than the original rumen contents, and PAIN tended (P =.08) to be lower in the preincubate for this fraction. Preincubation had no effect on the total PA or PAIN recovered in all three fractions. It was hypothesized that the high concentration of starch and maltose and the high dilution rate would select for proteolytic organisms. In addition, the dilution would remove non-microbial protein from the fluid. It appears that both hypotheses were false. Extraction in bufl’er, detergent or acetone Total PA was higher when treated with deoxycholate, than for the original pellet, but other treatments did not increase total PA (Table 7). Detergent may have altered protein conformation to increase digestion. Since the rumen does not contain detergent, it may not be a desired method to increase PA for in vitro studies. Mixing rumen fluid 120 precipitate with buffered media solubilized 18% of PA. Treatment with detergent or acetone solubilized 51 and 48 % of PA respectively. Activity per mg N increased 3.9 times in supernatant compared to original sample after treatment with detergent or acetone. These results demonstrate the potential of developing an in vitro system with enzyme extracted from pelleted rumen fluid. Gentle extraction with buffer may be advantageous because only cell wall proteases would be extracted and these are thought to be primarily responsible for degradation in vivo (Kopecny and Wallace, 1983). On the other hand, much more activity was recovered by extraction with acetone, so this enzyme is potentially more representative of whole rumen contents. Previous work has used enzyme recovered from acetone extraction for in vitro studies (Mahadevan, 1987). Treatment with acetone or detergent seemed more effective than preincubating, chilling, rinsing or blending of rumen contents at increasing proteolytic activity per unit protein of rumen preparations. 121 REFERENCES Broderick, G. A. (1987). Determination of protein degradation rates using a rumen in vitro system containing inhibitors of microbial nitrogen metabolism. Br. J Nutr. 58: 463- 475. Chamey, J. & Tomarelli, R. M. (1947). A colorimetric method for the determination of the proteolytic activity of duodenal juice. J. Biol. Chem. 171: 501-505. Craig, W. M., Hong B. J., Broderick G. A., & Bula, R. J. (1984). In vitro inoculum enriched with particle-associated microorganisms for determining rates of fiber digestion and protein degradation. J. Dairy Sci. 67: 2902 Furchtenicht, J. E. & Broderick, G. A. (1987). Effect of inoculum preparation and dietary energy on microbial numbers and rumen protein degradation activity. J. Dairy Sci. 70: 1404-1410. Goering, H. K. & Van Soest, P. J. (1970). Forage fiber analyses (apparatus, reagents, procedures, and some applications). Agric. Handbook No. 379. USDA-ARS. Washington, DC. Hungate, R. E. (1966). Chapter II: The rumen bacteria. The Rumen and Its Microbes. New York: Academic Press. Kohn, R. A. 1993. In vitro methods to determine protein degradation of feeds for ruminants. Ph.D. Thesis, Michigan State University, East Lansing, MI. Kohn, R. A. & M. S. Allen. (1993) An in vitro method to measure protein degradation of feeds using enzymes extracted from rumen contents. J. Dairy Sci. 76: Suppl. 1 Kopecny, J. & Wallace, R. J. (1982). Cellular location and some properties of proteolytic enzymes of rumen bacteria. Appl. Environ. Microbiol. 43: 1026-1033. Lin, K. W., Patterson, J. A. & Ladisch, R. R. (1985) Anaerobic Fermentation: microbes from ruminants. In: Enzyme and Microbial Technology 7: 98-107. Mahadevan, S., Erfle J. D., & Sauer, F. D. (1979). A colorimetric method for the determination of proteolytic degradation of feed proteins by rumen microorganisms. J. Anim. Sci. 48: 947—953. Mahadevan, S., Sauer F. D., & Erfle, J. D. (1987). Preparation of protease from mixed rumen microorganism and its use for the in vitro determination of the degradability of true protein in feedstuffs. Can. J . Anim. Sci. 67: 55. Neutze, S. A., Smith R. L., and Forbes, W. A. (1993). Application of an inhibitor in vitro method for estimating rumen degradation of feed protein. Animal Feed Science and Technology 40: 251-265. Pichard, G. R. (1977). In vitro rumen fermentation and the assessment of the nutritive value of forages. Ph.D. Diss., Cornell University, Ithaca, NY 122 SAS Institute Inc. (1991). Statistical Visualization for the Macintosh. JMP ver. 2.02. Cary, NC: SAS Institute Inc. Wallace, R. J. & Kopecny, J. (1983). Breakdown of diazotized proteins and synthetic substrates by rumen bacterial proteases. 123 Optical Density 4.. 2-1 600 700 800 900 Wavelength Fig. 1. Optical density of azocascin solution ( o ), centrifuged rumen fluid pellet ( D ), and centrifuged rumen fluid supernatant ( 9 ) at varying wavelengths of light. 124 Strained Rlnee Strained Rlnee Strained Rlnee Strained Rlnee Rumen oi Rumen of Rumen oi Rumen of Fluid Solid. Fluid Sollde Fluid Sollde Fluid Solide Blended Not Blended Blended Not Blended Chilled Not Chilled Azocasein degradation (mg degraded] hl ml) Fig. 2. Treatment of rumen fluid on proteolytic activity per m1 rumen fluid (E5 supernatant, I 20,000 x g pellet, I 400 x g pellet). Ruminal contents were chilled for 4 h or analyzed immediately, blended for l min or not, and analyzed directly after straining or solids were rinsed with an equal volume as rumen fluid of in vitro media and the rinse analyzed. The mean of three replicates are shown. Chilling and blending effects were not significant, strained > rinsed, 400 x g pellet > 20,000 x g pellet > supernatant, chilling x centrifugation and rinse x centrifugation interactions were significant, P = .01. 125 Strained Rinse Strained Rinse Strained Rinse Strained Rinse Rumen 0f Rumen of Rumen of Rumen of Fluid Solids Fluid Solids Fluid Solids Fluid Solids Blended Not Blended Blended Not Blended Chilled Not Chilled ’z‘6_ a: E 35 \ 8 .04 re a «1:3 '0 a) £2 5 :1 tr: '0 E m0 o 'o .E o en re 0 o N < Fig. 3. Treatment of rumen fluid on proteolytic activity per mg nitrogen in enzyme preparation. Ruminal contents were chilled for 4 h or analyzed immediately, blended for 1 min or not, and analyzed directly after straining or solids were rinsed with an equal volume as rumen fluid of in vitro media and the rinse analyzed. The mean of three replicates are shown. Chilling and blending effects were not significant, rinse > strained, P = .05. 126 Table I. Carbohydrate composition of media used for incubation of ruminal organism in continuous culture. Carbohydrate Amount (glL) Starch" 6.7 gIL Cellulose" 5.3 g/L Pectin 0.85 g/L Maltose 0.5 gIL Malic acid 0.2 g/L Sucrose 0.2 g/L Xylan 0.2 g/L Citric acid 0.12 g/L Cellobiose 0.1 g/L Glucose 0.1 g/L Succinic acid 0.08 E2 *Added separately in 2 parts at 12 h intervals. 127 Table 2. Treatment of rumen contents on proteolytic activity per ml rumen fluid (PA) and per mg nitrogen (PAW). Effect df PA (P) PAIN (P) Run (rep) 5 .0001 .0001 Chilling (C) l .09 NS Blending (B) 1 NS .01 C x B l .06 NS Error 1 15 NS NS Rinse vs. Fluid (R) 1 .0001 .03 R x C 1 .08 .03 R x B 1 NS NS R x C x B 1 .07 .05 Error 2 20 NS NS Centrifugation fraction (F) 2 .0001 NS F x C 2 .01 NS F x R 2 .001 .01 F x C x B 2 .06 .07 F x C x R 2 .0001 .08 F x B 2 NS NS F x B x R 2 NS NS F x C x B x R 2 NS NS Errorl(df)=erep(5)+Bxrep(5)+Cxerep(5). Error2(df)=erep(S)+R xerep(5)+Rxerep (5)+RxCxB xrep (5). Error3=allotherinteractions with rep. Error 1 was tested against Error 2 and Error 2 against Error 3. Effects rep, C, BandCwaeretested againstError 1. EffectsR,RxC,RxBandeCwaere tested against pooled Errors 1 and 2. Effect F and its interactions were tested against pooled Errors 1 to 3. NS = not significant at P> .1. Table 3. Mean proteolytic activity per ml rumen fluid (PA) and per mg N in rumen 128 fluid (PAW) . Main Effects PA PAIN Chilling Chilled 6.60 4.20 Not chilled 6.06 4. 2 6 Standard error of mean (11 = 12) .22 .1 1 Blending ** Blended 6.42 4.02 Not blended 6.30 4.50 Standard error of mean (n = 12) .22 .1 1 Fluid or rinse ** * Rumen fluid 3.66 4.08 Rinse of solids 2.64 4.74 Standard error of mean (n = 24) .13 .20 Centrifugation pellet ** 400 x g pellet 1.44 4.2 20,000 x g pellet .96 4.86 Supernatant .78 4.38 Standard error of mean (n = 48) .05 .03 Total activity and N recovered (sum of fluid and rinse for 3 centrifugation fractions) is shown for chilled, not chilled, blended and unblended treatments. Total activity and N of rumen fluid or rinse (sum of 3 centrifugation fractions) is shown for rumen fluid and rinse of solids. Mean activity in centrifugation fractions (for rumen fluid and rinse) is also indicated. *Indicates significant main effect at P = 05. **Indicates significant main effect at P = .01. 129 Table 4. Effect of treatment of rumen contents (chilling, rinsing and blending) on recovery of proteolytic activity per ml rumen fluid (PA) and per mg N (PA/N) in fractions separated by sequential centrifugation. PA (P) PAIN (P) Effect Pellet l Pellet 2 Supern. Pellet l Pellet 2 Supern. Chilling .001 .01 NS NS NS NS Blending NS .08 NS .001 .01 NS Chilling x blending NS NS .01 .001 NS .05 Rinse .01 .0001 .08 .01 NS .01 Rinse x chilling .0001 .01 .06 .0001 NS NS Rinse x blending .05 NS NS .05 NS NS R x C x B NS NS .08 .07 NS .07 Pellet 1 = 400 x g pellet, Pellet 2 = 20,000 x g pellet, Supern. = supernatant). Data analyzed in a split plot design according to Tab. 1 and description in text. 130 Table 5. Mean proteolytic activity per ml rumen fluid (PA) and per mg N in rumen fluid (PAflV) by centrifugation fraction. PA PA/N Effect pellet 1 pellet 2 supem. pellet 1 pellet 2 supem. Chilling ** ** Chilled 3.18 2.10 1.38 3.96 4.86 3.84 Not chilled 2.70 1.50 1.86 4.02 4.86 4.2 SEM (n = 12) .08 .09 .15 .09 .22 .31 Blending * Blended 2.94 1.8 1.68 3.78 4.38 4.02 Unblended 2.88 1.98 1.50 4.26 5.34 4.02 SEM (n = 12) .08 .09 .15 .09 .22 .31 Fluid or rinse ** ** * Fluid 1.5 1.26 0.96 3.84 4.98 3.36 Rinse 1.38 .6 .66 4.56 4.68 5.46 SEM (n = 24) .035 .044 .10 .14 .26 .52 Pellet 1 = 400 x g pellet, Pellet 2 = 20,000 x g pellet, Supern. = supernatant. Total activity and N recovered (sum of fluid and rinse) is shown for chilled, not chilled, blended and not blended treatments. Mean activity in for rumen fluid and rinse is also given. * Indicates significant main effect at P = .05. **Indicates significant main effect at P = .01. 13 1 Table 6. Efi'ect of preincubation in continuous culture at dilution rate of .07/h on proteolytic activity per ml rumen fluid (PA) and per mg N in rumen fluid (PAW) by cenmficgation fraction. PA (mglhlml) PAIN (mg/h/mg N) Treatment Pellet 1 Pellet 2 Supern. Total Pellet l Pellet 2 Supern. Total Effect * it * ¢ Preincubated 1.92 .54 .42 2.88 3.66 4.62 3.30 4.32 Rumen fluid 1.26 1.02 1.56 3.84 5.10 4.14 5.16 4.56 SEM (n=6) .14 .07 .25 .39 .36 .42 .61 .36 Pellet l = 400 x g pellet, Pellet 2 = 20,000 x g pellet, Supern. = supernatant. "' Indicates significant main effect at P = .05. ”Indicates significant main effect at P = .01. 132 Table 7. Efl'ect of solubilizing rumen fluid precipitate in reduced media, media with detergent, and acetone on proteolytic activity per ml rumen fluid (PA) and per mg N (PAW). PA (mg/hlml) PAIN (mghlmg N) Solvent Pellet Supern. Total Pellet Supern. Total Media 1.26ti .28c 1.5b 3.783 7.32b 4.1413 Detergent 1.02b 1.083 2.1a 3.63 15.12a 5.943 Acetone .72c 0.66b 1.36b 2.46b 15.183 4.08ID Untreated 4- - 1.36b -- - 3.9b SBM .031 .029 .064 .124 1.164 .237 After treatment in acetone, samples were centrifuged and pellet re-suspended in media. This and other treatments were centrifuged at 20,000 x g before determination of PA on azocasein and N. Different superscripts within a column indicate significant differences among treatments (P=.01). SEM = standard error of mean, 11 = 4. 133 CHAPTER 5: AN IN VITRO METHOD TO MEASURE PROTEIN DEGRADATION OF FEEDS USING ENZYMES EXTRACTED FROM RUMEN CONTENTS. Synapsis A method was developed to measure protein degradation of feeds using proteolytic enzymes extracted from rumen contents. Strained rumen fluid (1.01) was centrifuged and the pellet was extracted with butanol and acetone. Enzymes were then solubilized from the residue in water, and frozen at ~25°C for later use. Feed protein was degraded by incubating feeds with enzyme at 38°C. At the end of degradation, trichloroacetic acid (15 m1; 240 gIl) was added to precipitate undigested protein. The final enzyme preparation maintained 62% of the proteolytic activity on azocasein of the original rumen fluid (plus rinse of solids). When enzyme was incubated at 38°C, activity on azocascin decreased slightly for 12 hours and then increased. There was no difference between 3, 5 or 10 m1 of enzyme on degradation of soya bean meal or luceme hay protein at 6 or 16 h. This suggests that enzyme was supplied in excess for the duration of the incubation. Soya bean meal protein degraded at .15/h for the first 2 h and slowed to .01/h from 8 to 24 h. Similarly, luceme hay protein degraded at .06/h for the first 2 h and then slowed to .01/h. The proposed method is simple with low variation within treatment, and it uses enzymes from the rumen in concentrations that are maintained in excess of substrate for extended periods of incubation. Introduction Current systems to balance dairy cattle rations emphasize the need to consider ruminal protein degradation of feeds (V érité, 1979; Agricultural Research Council, 1984; 134 Madsen, 1985; National Research Council, 1989). Methods to evaluate ruminal protein degradation have limitations which restrict their use as has been reviewed (Nocek, 1988; Michalet—Doreau and Ouldbah, 1992). In vivo methods are time consuming and expensive, and the lack of adequate microbial and passage markers results in problems of accuracy and precision. In situ methods are the most commonly used procedures to measure protein degradation. Feeds are suspended in bags placed in the rumen of fistulated cows or sheep for differing durations of time. The residue left in the bag after in situ degradation is analyzed for nitrogen (N). Microbial contamination in the bag is occasionally determined by using a microbial marker such as nucleic acids or diamino pimelic acid (DAPA) which are associated with microbial matter but assumed to be digested from the original feed. In situ methods actually measure protein which does not pass out of the nylon bags in the rumen. Once a protein is solubilized or even reduced adequately in particle size to leave the bag it is considered degraded. Much of the protein of feeds leaves nylon bags before being completely degraded when incubated in vitro (Spencer et al. 1988). Various authors have developed methods to measure protein degradation that use proteolytic enzymes. Two of the most commonly used enzymes are ficin (Poos- Floyd et al. 1985) and endo-protease from S. teptomyces griseus (Krishnamoorthy et al. 1983). Ficin is a cysteine protease (Barrett, 1985) as are many of the proteases found in the rumen (Brock et al. 1982). However, some authors have criticized the use of plant-derived proteases to estimate degradation of protein sources by microbial proteases. On the other hand, S. griseus protease demonstrates maximal activity at pH 8 (Krishnamoorthy et al. 1983) which is not physiological. A newer method uses neutral protease of bacterial origin (Assoumani et al. 1992). Other in vitro methods use enzymes from ruminal organisms. While one approach is to incubate feeds with microbes in the presence of drugs that inhibit microbial growth (Broderick, 1987; 135 Neutze et al. 1993), another is to extract enzymes from ruminal organisms (Mahadevan et al. 1987). These enzymatic methods have two major limitations. The enzyme activity declines over time as the proteolytic enzymes degrade themselves (Broderick, 1987; Mahadevan et al. 1987). Therefore, these methods are only appropriate for short-term degradation (1-2 b). However, feeds may be comprised of multiple protein fractions that degrade at different rates. The total feed protein may initially degrade fast and then slow due to disappearance of the fast fraction. Because of this, methods are required that measure protein degradation over extended periods of time. A second limitation is that first order degradation kinetics do not apply to these systems because the level of enzyme activity limits the degradation velocity (Mahadevan et al. 1987; Broderick and Clayton, 1992). Though these methods are used to detect some relative differences among feeds, they have not been adequate to determine fractional degradation rates because rate of degradation under enzyme-limiting conditions depends on feed sample size, types of enzymes and level of activity. The objective of this work was to develop and validate a method to measure protein degradation using enzymes extracted from rumen contents, and to maintain activity over time, so that enzyme activity would be supplied in excess for extended periods of incubation. MATERIALS AND METHODS Solutions and feeds Buffered media used in these experiments (except where indicated otherwise) was modified from Goering and Van Soest (1970) by replacing NH4HCO3 with an equimolar concentration of NaHCO3. The media was diffused with C02 gas before use until pH stabilized at 6.8. At this time, media was reduced by addition of 1.2 gIl of 136 each Na28.(1120)9 and L-cysteine.HCl.I-120 (Sigma Chemical Co., St. Louis, MO, USA). The media was maintained under 1 atm C02 to prevent oxidation or an increase in pH due to loss of C02. . Feeds used in development of procedures and their respective percentage crude protein content on a dry matter basis were: solvent extracted soya bean (Glycine max) meal - 48.0, luceme (Medicago sativa) hay at one half bloom -- 19.1, dried and ground maize (Zea mays L.) grain -- 10.8, and dried meat and bone meal -- 55.2. Each feed was ground through a 1 mm screen of a Wiley mill. Butanol-acetone extraction The method used to extract enzyme from rumen contents was modified from that of Mahadevan et al. (1987) based on observations recorded elsewhere [Appendix] (Kohn, 1993). Rumen fluid (5 l) was collected from 2 fistulated cows 24 h after feeding. The cows were fed a diet based on luceme hay and maize silage twice daily. Rumen fluid was blended for 1 min and strained through cheese cloth, though results of another study showed that rumen contents should not be blended as blending decreased proteolytic activity per mg N [Chapter 4] (Kohn and Allen, Submitted). The solids were twice rinsed with reduced in vitro media (1.25 l), and this was added to the strained rumen fluid. This mixture was stored at 38°C under C02 until it could be processed (0-5 h). The following extraction procedure was repeated up to six times within a day. Rumen fluid (1.0 l) was centrifuged for 30 min at 4°C and 13,000 x g. It was decanted with care taken to keep loose sediment with pellet. Water was added to pellets to bring the total volume up to 250 ml. Equal volume cold (4°C) butanol was added to pellets and mixed for 1 h using a magnetic stir bar over an ice-water bath. After an hour, this mixture was added intermittently for 45 min to acetone (1.01) which was also stirred with a magnetic bar over an ice-water bath. The acetone was stirred for an additional 137 15 min and then filtered through Whatman # 54 filter paper in a Buchner funnel with suction. The filtrand was rinsed with two 500 m1 aliquots of cold acetone and dried in the funnel. Each pellet was then stored overnight in a refrigerated desiccator. The following day, pellets were each resolubilized in 150 ml distilled water in centrifuge bottles. The bottles were shaken vigorously to wet the sample, and were placed in a shaking ice-water bath for one hour. The fluid was centrifuged for 30 min at 26,000 x g and 4°C, and decanted while keeping loose sediment with supernatant. The loose sediment was removed later by centrifugation before using the rumen extract. The supernatant contained enzyme activity and was frozen at -25°C. After decanting and saving the supernatant, an additional 100 m1 of water (4°C) was added to the pellet, the bottle shaken again for 1 h and centrifugation and decanting repeated. Feed protein degradation Feed samples containing 8 mg nitrogen were weighed into 50 ml plastic centrifuge tubes. Tubes were placed under C02 before adding in vitro media (25 m1). Reduced media was added because previous work showed that proteins may precipitate in the presence non—reduced media, and this could alter degradation (Kohn and Allen, 1992). Enzyme preparation was thawed, centrifuged (30 min, 26,000 x g), and decanted with exclusion of loose sediment before adding supernatant to tubes (6 ml) and incubating at 38°C under C02. Trichloroacetic acid ( 15 ml, 24%) was added to end the degradation and precipitate undigested protein. Samples were centrifuged at 1500 x g for 20 min and the supernatant discarded. A modified Kjeldahl method (Hach et al. 1985) was used to determine nitrogen content in the pellet, and undegraded protein was determined as nitrogen times 6.25. 138 Reducing agents Proteolytic activity was determined on azocasein as described [Chapter 4] (Kohn and Allen, Submitted) except for addition of reducing agents. A control with no reduction was compared to two levels of cysteine.HCl, cystine, sodium sulphide (.6 g/l or 1.2 g/l) and dithiothreitol (2 or 4 mM). The experiment was repeated on two different days with two replicates of each treatment on each day. Effect of reduction on soya bean meal degradation was determined. Soya bean meal was degraded for 16 h with extracted enzyme preparation as described previously. Treatments included: 1. media that was neutralized with HCl but not saturated with C02 (aerobic), 2. media that was equilibrated with C02 but not reduced (anaerobic), 3. media that was reduced with 1.2 g/l sodium sulphide and incubated with constant perfusion of 1 atm C02 (open), and 4. media was reduced with 1.2 g/l sodium sulphide and maintained in stoppered centrifuge tubes under C02 so that sulphide could not escape (closed). Results were compared using the Student t test. All statistics in these studies were performed with J MP (SAS Inst. 1992). Cellulase and amylase One microliter alpha-amylase (T ermamyl, Novo Nordisk Bioindustn'als, Inc., Danbury, CT, USA) was added to maize grain samples incubated with enzyme as described. This amount of amylase was estimated to be adequate to maximize fractional degradation rate of starch. Commercially-produced cellulase from Trichiderma virde (Sigma Chemical Co., St. Louis, MO, USA) was added to samples containing protease preparation. There were no differences in protein degradation with cellulase at 16 h because the cellulase was minimally active at pH 6.8, and was rapidly degraded by protease. Thereafter, a two-step procedure was used for degradation with cellulase. 139 . Degradation with cellulase was adapted from Jones and Hayward (1973). A citrate-phosphate buffer was prepared (.133 M Citric acid, .047 M dibasic sodium phosphate and .065 M sodium chloride), equilibrated with 1 atm C02, and adjusted to pH 4.6 with sodium hydroxide or hydrochloric acid. In the first step, luceme hay was incubated with 10 ml citrate-phosphate buffer with cellulase from T. virde (50 rug/sample; Sigma Chemical Co., St. Louis, MO). In the second step, 15 ml bicarbonate buffer was added to samples to raise the pH to 6.75 (+I- .01). Bicarbonate buffer (.12 M sodium bicarbonate, .21 ml micro-mineral solution (Goering and Van Soest, 1970), 1.0 mM magnesium chloride) was prepared by equilibrating with 1 atm gaseous C02, reducing with 2 g/l of each cysteine.HCl.H20 and Na28.9H20 and then adding 5 N KOH (9 mill). Enzyme preparation extracted from rumen contents (6 ml) was added to neutralized samples. Several variations were compared to determine the effect of cellulase pretreatment on protein degradation by extracted enzyme. Effect of cellulase pretreatment for 4 or 16 h was determined with or without degradation with rumen enzyme for 4 or 16 h respectively. These were compared to samples with no pretreatment or that were soaked in pH 4.6 buffer without cellulase for 4 or 16 h before degradation with rumen enzyme. Nitrogen was determined after precipitation with TCA as described, and dry matter and nitrogen were determined on additional samples that were filtered with suction through Whatman # 54 filter paper after degradation without addition of TCA. Each treatment was replicated 3 times. Results were analyzed statistically using AN OVA to compare treatment effects by time and time effects by treatment for each response variable. Proteolytic activity over time The enzyme preparation (.25 mll tube) was dispensed to 15 ml centrifuge tubes. Reduced in vitro media (.75 ml) was added under C02, and samples were stoppered and frozen at -25°C. Samples were removed from the freezer and incubated at 38°C for 140 0, l, 2, 3, 4, 8, 12, 16, or 24 h before addition of reduced azocascin solution (12 mg, 1 m1). Samples were removed from the freezer at different times and all treatments were combined with azocascin at the same time. Two-hour azocascin degradation was completed twice using different enzyme preparations with two replicates per run as described [Chapter 4] (Kohn and Allen, Submitted). Efl'ect of enzyme level Soya bean meal and luceme hay were subjected to incubation with 0, l, 3, 5 or 10 ml of enzyme preparation. Protein degradation at 16 h was determined as previously described. Three replicates of each treatment were used. A different experiment was to test whether enzyme activity remained in excess for the duration of the experiment. Soya bean meal was incubated for 0, 8 and 16 h with 5 and 10 m1 enzyme. An additional treatment was soya bean meal incubated with 5m] enzyme for 8 h, and then 5 ml more enzyme added, followed by further incubation for 8 h. Feed protein degradation was determined as described previously except enzyme preparation was not centrifuged before use because importance of this step had not been determined. The experiment was repeated twice. Comparison of feeds and variation of method Degradation of soya bean meal and luceme hay was determined at 0, 2, 4, 8, 16, 24, 32 and 48 h using the method described for protein degradation of feeds. The experiment was repeated twice with rumen fluid from different days. Rate of protein degradation was determined as the natural log of the fraction degraded in a time interval divided by that time interval. Four feeds described previously (soya bean meal, luceme hay, maize grain and meat and bone meal) were compared for protein degradation using the final method described previously. Each feed was digested for 6 and 16 h. The experiment was 141 repeated three times using rumen fluid extracted on different days in order to evaluate day to dayvariation. Each treatment was replicated three times each day in order to evaluate within day and day x treatment variation. RESULTS AND DISCUSSION Butanol-acetone extraction After extraction with butanol and acetone, 62 % (SD. = 10, n = 5) of the rumen fluid proteolytic activity on azocascin was solubilized in water. Activity was either recovered in the first (75 %, SD. = 5, n = 17) or second extraction with water (25 %, SD. = 5, n = 17), with little recovery in the third extraction. Therefore, the standard procedure was to use only two extractions with water. A second centrifugation at 26,000 x g on the extracted rumen enzyme reduced TCA precipitable nitrogen in the enzyme preparation below detectable levels without any reduction in proteolytic activity. This enzyme preparation therefore enabled determination of feed protein degradation without interference in measurement from proteins in the enzyme preparation. A different study showed that 50% of proteolytic activity could be recovered in one extraction with buffer after extracting rumen fluid solids with acetone [Chapter 4] (Kohn and Allen, submitted). Extraction in acetone is somewhat easier than butanol- acetone but does not remove pigments from the final solution, and therefore would interfere with colorimeuic measurement of protein. However, it could be sufficient for determinations based on N. A different method enabled recovery of 30% of proteolytic activity from rumen fluid after extraction in butanol-acetone (Mahadevan et al. 1987). One reason the method used to extract protease in the present study resulted in higher recovery of activity compared to the previous method is because previously enzyme preparation was filtered which results in a loss of some activity. Contamination from 142 protein in the enzyme preparation was a limitation of the previous work. This problem was eliminated with the additional centrifugation step, and final separation with TCA. The proposed method of extraction is also less labor intensive than the previous method. Reducing agents Since some effects of addition of chemical reducing agents on degradation of azocascin were inconsistent across runs, each run was analyzed separately. In both runs, addition of cysteine.HCLH20 increased proteolytic activity. Comparison of level of cysteine using student t test showed that 1.2 mg/ml resulted in more protein degradation than .6 mg/ml in both runs. Sodium sulphide addition resulted in a slight increase in PA over control, which was significant only for the lower level in one run. Sulfide (H28) is easily lost from the system by diffusion into the C02 head space of tubes. This may explain its inconsistent behavior. Dithiothreitol (DTT) and cystine had no effect on protein degradation in either run. Cystine was added as a negative control in that it is not a reducing agent but is in fact the oxidized form of cysteine.HCl. The dramatic results shown in Table 1 indicate that the enzyme extracted from the rumen clearly benefits from addition of reducing agents. Simply excluding oxygen from the media did not adequately reduce it. Additional experiments demonstrated the importance of reducing agents on luceme hay degradation, and there was no difference between sulphide, cysteine or a combination thereof on degradation of soya bean meal protein [Appendix] (Kohn, 1993). Based on these results, 1.2 gll of each reducing agent (Na28 and cysteine.HCL) were added to media after equilibrating with 1 atm C02, and media was maintained under a 1 atm C02 environment with care taken to limit the exchange of gases to prevent evolution of H28. 143 Proteolytic activity over time The 2-h degradation of azocascin (proteolytic activity) decreased when rumen fluid was first incubated without azocascin. This decrease was at a rate of .049Ih for 0 to 12 h (Figure 1). Surprisingly, the proteolytic activity at 24 h was much greater than the initial level. When this experiment was repeated to verify these results, proteolytic activity was variable over time and degradation in activity was not significant. Again degradation at 24 h was as high as in the data shown. It appears that the enzymes may be auto-activated by incubation in vitro (Lehninger, 1975). The slow rate of degradation of enzyme (or no degradation) and the possibility for auto-activation is promising in that enzyme preparations may be viable for degradation of feeds over extended periods of time. Cellulase and amylase Addition of amylase had no effect on protein degradation of dried maize grain at 16 h. The volume of maize grain was visibly reduced in all treatments, and the residue after degradation appeared to be a gummy substance. Enzymes extracted from rumen contents using a similar method demonstrated the presence of amylolytic activity (Mahadevan et al. 1987). Previous work has demonstrated that addition of amylase increases protein degradation due to treatment with commercially available protease (Assoumani et al. 1992). In the present study, there may have been no effect of amylase addition because starch was adequately removed by extracted enzyme that it did not interfere with protein degradation. Cellulase pretreatment decreased protein degradation of luceme hay with protease slightly at 4 h, and increased it dramatically at 16 h (Table 2). The decrease at 4 h may have been caused by an incorrect estimate of feed residual protein due to contamination by cellulase. Blank samples containing cellulase and enzyme but no 144 feed were incubated for 4 or 16 h, and the residual protein subtracted from those with feed. The contaminating protein may have degraded more slowly with feed samples than with blanks due to the additional protein occupying sites on the enzymes. By 16 h, cellulase protein was not detectable in the blanks. The dramatic increase in protein degradation after 16 h in protease when preheating with cellulase indicates that fiber may interfere with protein degradation especially during long-term degradation of forages. There was a slight degradation of protease over time when incubated only with cellulase. This could be due to proteolytic activity of cellulase, contaminants in the enzyme preparation, or the plant's own proteolytic enzymes. The fact that pretreatment in pH 4.7 buffer before proteolytic degradation does not increase protein degradation over protease alone suggests that the increased activity with cellulase is due to the cellulase and not to presoaking. Nitrogen solubility in buffer (pH 6.8) determined by filtering was greater than nitrogen solubility in TCA across treatments (Tables 2 to 3). This shows that T CA precipitation would be expected to provide a lesser value for protein degradation than in vitro or in situ methods that rely on solubility. Correlation between solubility of nitrogen in TCA and buffer was .73 and between nitrogen solubility in TCA and DM solubility was .66. The highest correlation was between DM and nitrogen solubility in buffer (.85). Dry matter (DM) degradation was greatest for samples pretreated with cellulase followed by protease degradation compared to just cellulase 0r protease (Table 4). Though each enzyme preparation resulted in significant DM disappearance, the combination was most effective. These results suggest that a substantial amount of protein in forages may be protected from degradation by fiber. Addition of cellulase may be important to the measurement of the long-term protein degradation of forages. 145 Efl'ect of enzyme level There was no difference in degradation of soya bean meal or luceme hay at 6 or 16 h when 3, 5 or 10 ml of enzyme preparation was used. There was a reduction in protein degradation when 1 ml enzyme was used compared to 3 ml. Subsequent experiments using this enzyme preparation used 6 ml of enzyme in order to use at least twice the amount needed to supply enzyme in excess. Table 5 shows the results of adding additional enzyme at 8 h to soya bean meal that had already been degraded for 8 h. There was a slight decrease in protein degradation by adding half the enzyme after 8 h degradation compared to adding all the enzyme initially. There was no difference between adding 5 ml or 10 ml of enzyme for degradation at either 8 or 16 b. If enzyme activity decreases overtime and is limiting to degradation of feeds, we would expect that addition of 5 ml enzyme after 8 h degradation would result in greater degradation at 16 h than addition of all enzyme at the start. Since this increase was not observed, it appears that enzyme activity was in excess for the duration of the incubation. The lower degradation at 6 and 16 h as compared to other experiments in this paper appears to have resulted from an underestimation of crude protein in blanks, since blank nitrogen was much higher because the enzyme preparation was not centrifuged before use. Another experiment demonstrated the importance of centrifuging the enzyme preparation before use on measurement of protein degradation [Appendix] (Kohn, 1993). Comparison of feeds and variation of method Protein degradation of soya bean meal and luceme hay averaged .15/h and .06/h respectively for the first 2 h of incubation (Figure 2). These rates slowed to .01/h from 8 to 16 h. These results emphasize the importance of long-term degradation of feeds to determine protein degradation. Prediction of protein degradation at extended periods of time from initial rates may overestimate protein degradation. 146 The results of degrading four protein sources using enzyme are shown in table 6. There was no difference in enzyme preparation obtained on different days for either point in time. However, there was a significant enzyme x run interaction at 16 h (P = .02). When effect of enzyme was determined by ANOVA on each feed, the enzyme effect was only significant for dried maize grain at 16 h (P = .04). Average degradation of maize grain at 16 h was 20.3, 23.8 and 28.8 % for each the three runs. Since maize grain was lowest in percentage protein the greatest quantity was used on a dry matter basis (since an equivalent amount of N was used for each feed). This may have resulted in the lower .degradation of maize than the other feeds and increased variation from run to run for maize. Overall, the low variation within and across runs is promising. The actual degradation of each feed protein was less than expected, especially at 16 h. The fact that these values may be lower than has been determined for similar feeds in situ (National Research Council, 1989) is not surprising because the in situ procedure would overestimate protein degradation as particles leave the bag prior to complete digestion. The degradation in the present study may also be lower than that observed for feeds degraded with rurrrinal enzyme in the presence of inhibitors (Broderick, 1987). The inhibitor system uses only short-term degradation with extrapolation out to longer time points. The short-term (2 h) degradation of luceme hay and soya bean meal was similar in the present study to that measured using the inhibitor system (Broderick, 1987). These same feeds were also degraded in vitro in a system employing sulphur-35 (358) to measure microbial growth [Chapter 3] (Kohn, 1993). The long-term degradation using the 353 system was higher for all feeds than in the present work. The system using 358 or other markers may also overestimate protein degradation because the free bacteria fraction separated with sequential centrifugation may be contaminated with feed protein from chloroplasts or mitochondria (Pichard, 1977). The fact that cellulase was not present in the enzyme preparation may explain 147 additional decrease for luceme. Other enzymes may also have been selected against or inactivated by extraction, or enzyme associations with microbes, may influence the degradation of some proteins. Enzymes that are ATP-dependent would be inactive in extracted preparations, and such proteases have been identified in E. coli (Matin, 1992). CONCLUSION Extraction of enzymes from rumen fluid provides a means to concentrate enzyme activity for degradation of feeds in vitro. This enzyme preparation seems to maintain its activity for extended incubations (> 16 h), but degradation of feeds slows markedly with longer incubation. Interaction of cellulase with protein degradation indicates that protein degradation may require more than the presence of protease for degradation Further research is needed to characterize this enzyme activity in the crude extract on different types of feed proteins, and other structural components that may interfere with protein degradation. 148 REFERENCES Agricultural Research Council (1984). The Nutrient Requirements of Ruminant Livestock, Suppl. 1. Slough: Commonwealth Agriculture Bureaux. Assoumani, M. B., Vedeau F., Jacquot L., and Sniffen, C. J. (1992). Refinement of an enzymatic method for estimating the theoretical degradability of proteins in feedstuffs for ruminants. Anim. Feed Sci. Technol. 39: 3-4 357-368. Barrett, A. J. (1985). Chapter 1: The classes of proteolytic enzymes. Plant Proteolytic Enzymes. 1: 1-16. Brock, F. M., Forsberg C. W., and Buchanan-Smith, J. G. (1982). Proteolytic activity of rumen microorganisms and effects of proteinase inhibitors. Appl. Environ. Microbiol. 44: $61-$69. Broderick, G. A. (1987). Determination of protein degradation rates using a rumen in vitro system containing inhibitors of microbial nitrogen metabolism. Br. J. Nutr. 58: 463-475. Broderick, G. A. and Clayton, M. K. (1992). Rumen protein degradation rates estimated by non- linear regression analysis of Michaelis—Menten in vitro data. Br. J. Nutr. 67: 27 -42. Goering, H. K. and Van Soest, P. J. (1970). Forage fiber analyses (apparatus, reagents, procedures, and some applications). Agric. Handbook No. 379. USDA-ARS, Washington, DC. Hach, C. H., Brayton S. V., and Kopelove, A. B. (1985). A powerful Kjeldahl nitrogen method using peroxymonosulfuric acid. J. Agric. Food Chem. 33: 1117 . Jones, D. I. H. and Hayward, M. V. (1973). A cellulase digestion technique for predicting the dry matter digestibility of grasses. J. Sci. Food Agric. 24: 1419- 1426. Kohn, R. A. (1993). In vitro methods to determine protein degradation of feeds for ruminants. Ph.D. Diss., Michigan State University, East Lansing, MI. Kohn, R. A. and Allen, M. S. (1992). Storage of fresh and ensiled forages by freezing affects fibre and crude protein fractions. J. Sci. Food Agric. 58: 215-220. Kohn, R. A. & M. S. Allen. Enrichment of proteolytic activity in preparations from the rumen for in vitro studies. Br. J. Nutr. Submitted. Krishnamoorthy, U., Sniffen C. J., Stern M. D., and Van Soest, P. J. (1983). Evaluation of a mathematical model of rumen digestion and an in vitro simulation of rumen proteolysis to estimate the rumen-undegraded nitrogen content of feedstuffs. Br. J. Nutr. 50: SSS-568. Lehninger, A. L. (1975). Biochemistry. The Molecular Basis of Cell Structure and Function. 2nd edition. Worth Publishers, Inc. New York. 149 Madsen, J. (1985). Protein evaluation for ruminants. Acta Agric. Scand. Suppl. 25: 9- Mahadevan, S., Sauer F. D., and Erfle, J. D. (1987). Preparation of protease from mixed rumen microorganism and its use for the in vitro determination of the degradability of true protein in feedstuffs. Can. .1. Anim. Sci. 67: 55. Matin, A. (1992). Physiology, molecular biology and applications of the bacterial starvation response. Journal of Applied Bacteriology 73: S49—S57. Michalet-Doreau, B. and Ouldbah, M. Y. (1992). In vitro and in sacco methods for the estimation of dietary nitrogen degradability in the rumen - a review. Animal Feed Science and Technology 40: 57-86. National Research Council (1989). Nutrient Requirements of Dairy Cattle. 6th ed., Washington, DC: National Academy Press. Neutze, S. A., Smith R. L., and Forbes, W. A. (1993). Application of an inhibitor in vitro method for estimating rumen degradation of feed protein. Animal Feed Science and Technology 40: 251-265. Pichard, G. R. (1976). In vitro rumen fermentation and the assessment of the nutritive value of forages. Ph.D. Thesis, Cornell University, Ithaca, NY. Poos-Floyd, M., Klopfenstein T., and Britten, R. A. (1985). Evaluation of laboratory techniques for predicting ruminal protein degradation. J. Dairy Sci. 68: 829-839. Spencer, D., Higgins T. J. V., Freer M., Dove H., and Coombe, J. B. (1988). Monitoring the fate of dietary proteins in rumen fluid using gel electrophoresis. Br. J. Nutr. 60: 241—247. Vérité, R. (1979). Appreciation of the nitrogen value of feeds for ruminants. Pages 87- 96in Standardization of Pages Analytical Methodology for Feeds. Eds. W. J. Pigden, C. C. Balch and M. Graham. .nfl E .2 an 55. OD 865 .E'E' 5932 O\ 000 at; Q a 0 I I I I I 0 5 10 15 20 25 Hours pre-incubated Figure 1. Effect of incubating rumen enzyme preparation at 38°C on proteolytic activity on azocascin. Error bars indicate standard error of the mean for points indicated; 11 = 2. A [ § 8'0 "U '8 1 £3 60. 00 d.) "U :1 s 40. .E 8. g 20. 0 l I 1 I I 0 10 20 30 40 50 Time (h) Figure. 2. Degradation of soya bean meal (—- El 4) and luceme hay (-— O --) with enzyme extracted from rumen fluid. Standard error of the mean for soya bean meal = 1.9 and for luceme hay = 2.6; n = 2. 152 Table 1. Efi‘ect of reduction of in vitro media on soya bean meal (SBM) degradation at 16 h in enzyme extracted from the rumen. Media SBM degradation (%) Aerobic 19.4a Anaerobic (C02) 21.5a Reduced (open) 48.0b Reduced (closed) 45.4b Media was reduced with 1.2 g/l Nazs and maintained under diffusion of C02 (open) or stoppered under C02 (closed). aDifferent superscripts indicate differences among means, it = 2, P < .05). 153 Table 2. Solubility of nitrogen from lucerne hay in trichloroacetic acid as afi‘ected by incubation with protease extracted from the rumen (P), commercial cellulose (C), soaking in pH 4. 7 bufi‘er before P, and C before P. Time effect (P) Treatment 0h 4h 16h 0v.4h 4v.16h Protease (P) 27.8 37.3til 42.4at .0005 .01 Cellulase (C) 27.8 275') 32.913 NS .004 Buffer, P 27.8 37.03 40.94b .003 .09 C, P 27.3 31.2e 51.2e NS .002 SEM .89 1.09 2.47 -- ..- SBM = standard error of mean, 11 = 3. Different superscripts within a column indicate differences among means (P < .05). 154 Table 3. Solubility of nitrogen from luceme hay in bufi‘er (pH 6. 8) as affected by incubation with protease extracted from the rumen (P), commercial cellulose (C), soaking in pH 4. 7 bufier before P, and C before P. Time effect (P) Treatment 0h 4h 16h 0v. 4h 4v.16h Protease (P) 38.9 45.5a 52.221?!) .004 .008 Cellulase (C) 38.9 37.2b 43.43 NS NS Buffer, P 38.9 46.13 seat .0002 .001 C, P 38.9 58.90 629‘) .0001 .08 SEM ' .59 .55 2.12 m --- SEM = standard error of mean, 11 = 3. Different superscripts within a column indicate differences among means (P < .05). 155 Table 4. Solubility of dry matter (DM ) from luceme hay in bufier (pH 6. 8) as affected by incubation with protease extracted from the rumen (P), commercial cellulase (C), soaking in pH 4. 7 buffer before P, and C before P. Time effect (P) Treatment 0h 4h 16h 0v. 4h 4v. 16h Protease (P) 24.7 29.44 35.2a .0001 .0001 Cellulase (C) 24.7 35.5b 42.4b .0003 .003 Buffer, P 24.7 29.9a 33.1a .0002 .002 c, P 24.7 46.6c 53.1c .0001 .0001 SEM .099 .384 .61 --- --- SEM = standard error of mean, it = 3. Different superscripts within a column indicate differences among means (P < .05). 156 Table 5. Efl‘ects of enzyme level and addition of enzyme after 8 h on soya bean meal degradation at 8 and 16 h. Enzyme added (ml) Protein degraded 0 h 8 h 0 h 8 h 16 h 10 0 1.3 29.7 39.5‘1 5 0 8.1 26.2 39.8a 5 s -- 36.49 SEM -- 2.1 2.2 .64 SEM = standard error of the mean, 11: 2. Different superscripts within a column indicate differences among means (P < .05). 157 Table 6. Average crude protein degradation (percentage of original) after 0, 6 or 16 h incubation with enzyme extracted from the rumen. Degradation (%) . Feed 0 h 6 h 16 h Soya bean meal 5_7a 37.9a 47_3a Luceme hay 24.1c 38.2a 41.7b Meat and bone meal 15.0b 36.68 424*) Maize grain (dried, ground) 13.6b 24.8b 24.3c Standard error 1.006 1.073 .870 of mean (n = 9) Different superscripts within a column indicate differences among means (P < .05). ——__._ 3.5. Lira-v 1r : -' :—£»‘-f."-'." ”gang... _ , - .-_-_ —- -- 158 CHAPTER 6: EFFECT OF PLANT MATURITY AND PRESERVATION METHOD ON IN VITRO PROTEIN DEGRADATION OF F ORAGES ABSTRACT The influence of maturity and method of conservation on protein degradation was determined on four different forage species. Alfalfa, smooth bromegrass and reed canary grass were harvested at three maturity levels and whole plant corn was harvested at two maturity levels. Samples of each forage were freeze dried, or prewilted and ensiled in mini-silos at two DM contents. Additional samples of all forages except corn were field dried to hay. Ground sample was incubated for O, 2 and 24 h with crude enzyme extract from rumen contents. Degraded protein as a percentage of total crude protein was determined as the protein soluble in trichloroacetic acid (80 g/L) after degradation. Increased maturity resulted in lower protein degradation for alfalfa, bromegrass and canarygrass. Though ensiling increased protein degradation compared to dried forage, there was no consistent effect of silage DM content. Freeze dried material generally had less degraded protein than hay, but bromegrass protein degradation at 24 h was lower in hay than freeze dried samples. Protein degradation of forages was highly variable and depended on plant maturity and conservation method. (Key words: protein degradation, forages, maturity, hay, silage) Abbreviation key: TCA = Trichloroacetic acid 159 INTRODUCTTON Currently used systems to balance dairy cattle rations emphasize the need to consider ruminal protein degradation (1, 20). Forages generally have low crude protein content compared to protein supplements. However, since forages are the main ingredient of most dairy rations, they supply a significant portion of the total crude protein content of the diet. There has been little research on the protein degradation properties of forages. Since various methods to determine protein degradation have limitations (21), the few results that are available require confirmation using an alternative means to measure protein degradation. As plants mature, different proteins may be required to meet the needs of changing plant metabolism and structure. Conservation of forages may also result in changing profiles and tertiary structure of plant proteins. These changes may alter the proteolytic degradation of forage protein. The objective of this study was to determine effects of plant maturity and forage preservation method on in vitro protein degradation of four forage species. METHODS Alfalfa (Medicago sativa L.) was harvested in bud stage (Early), one tenth bloom (Mid), and full bloom (Late). Smooth bromegrass (Bromus inermis Leyss) and reed canarygrass (Phalaris arundinacea L) were harvested in boot (Early), early reproductive (Mid) and seeded (Late) stages. Whole plant corn (Zea mays L.) was harvested at half milk line (Early) and black layer formation (Late). Perennial forages were harvested in the first cutting of 1989 from randomized university plots that were established the previous spring. Corn was harvested from a different university plot. After harvesting withva hand sickle, each forage was mixed and then divided into parts to provide three replicates of each treatment. Each forage was preserved four different ways (three ways for corn) at each maturity level. Forage samples were frozen _" "' “FR" *4.“ 2.134e-15: - t 160 at -25°C, and then freeze dried with care taken to avoid thawing of samples, or they were prewilted for silage or hay (except for com). Hay samples were dried outdoors on pavement to prevent leaf loss in the field. Alfalfa hay at mid maturity was moved indoors and drying completed in front of a forced air heater (30°C) to prevent damage due to out-door precipitation. Samples of each forage were prewilted on pavement to 30-35% and 40-48% dry matter content before ensiling in mini silos (2.5 L) made from polyethylene tubing. Forages were coarsely chopped in a 24-knife hammermill, and packed into mini silos to a density of .8 kg/L. The moist silage from late corn was made by direct addition of water after chopping but before ensiling. The drier silage from early corn was made by prewiltin g indoors in front of a forced air heater (30°C) before chopping. After ensiling for at least 60 d, silage samples were frozen, and freeze dried. Crude dry matter content was determined from the difference in mass of original and freeze dried samples. Dried samples were ground through a 5 mm and then a 1 mm screen of a Wiley mill. Dry matter content of ground and dried forages was determined by drying at 100°C for 16 h (6), and this was multiplied by crude dry matter for final dry matter determination. All other analyses were determined on freeze dried and ground samples, and are expressed on a dry matter basis. Samples of each forage were ashed at 500°C for 5 h (6). Fiber fractions (NDF, ADF, lignin were determined sequentially (6)with modification of the NDF procedure by the substitution of triethylene glycol for 2- ethoxyethanol (18 g/kg) and the omission of decahydronaphthalene and sodium sulfite (24). Crude protein was determined as Kjeldahl nitrogen (8) x 6.25. Crude protein _~;.‘.A. n . degradation was determined after 0, 2 and 24 h incubation with enzymes extracted from rumen contents (11). Degraded crude protein was considered to be that soluble in trichloroacetic acid (TCA) at a final concentration of 80 g/L. 161 Statistics were performed using JMP (26). Each forage species was analyzed separately due to significant interaction effects of species with maturity and preservation method. Maturity, preservation method and the interaction effects were determined using AN OVA at each length of incubation. Further elucidation of main effects was made using the following specific contrasts: early vs. mid maturity, mid vs. late maturity, freeze dried and field dried vs. ensiled forages, freeze dried vs. field dried, and the wetter silage vs. the drier. RESULTS AND DISCUSSION Forage analysis results are presented for each forage species and maturity level (Table 1). Effects of maturity were as expected for each component shown (20). Crude protein, fiber, lignin and ash were slightly higher in silage and hay compared to freeze dried samples (P < .05, data not shown). These effects would have resulted from respiration of other feed components in samples during wilting or ensiling (7, 23). Alfalfa and bromegrass that were prewilted to a higher DM content before ensiling resulted in greater ash content compared to silage preserved at lower DM (P < .05). This indicates that lower silage DM resulted in reduced respiration, as would be expected from previous research (25). Alfalfa crude protein was degraded more than that of other forages regardless of length of incubation with enzyme (Table 2). Cool season grasses (bromegrass and canarygrass) had higher protein degradation than corn. Previous research on protein degradation of forages in situ confirms these relative differences (4, 9, 10 12, 18). A previous production trial showed that milk production increased more due to supplementation with unde graded protein when dairy cattle were fed diets based on alfalfa silage instead of corn silage (30). This effect was likely due to the deficiency of undegraded protein on the alfalfa silage diet. 162 Advanced maturity resulted in lower degradation of CP for alfalfa and bromegrass samples regardless of length of incubation (Table 2). The overall main effect of maturity on canarygrass CP degradation at 24h was significant (P < .05), though individual contrasts were not. There was no maturity effect on corn CP degradation. In previous research, advanced maturity was associated with lower disappearance of CP from nylon bags incubated in situ (9, 16, 29). However, maturity effects were not demonstrated with grass and legume silage studied in vivo (13). Ensiled forages were consistently higher than dried forages in CP degradation (Table 3), as has been demonstrated previously (5, 14, 22). Silage DM affected grass CP that was soluble in TCA before incubation with enzyme in vitro (0 h), but there was no effect of silage DM on CP degradation after 2 or 24 h incubation with enzyme. The initial increased CP degradation in drier silages of bromegrass may have resulted from proteolysis during the wilting process (15). However, drier silages of canarygrass and corn had lower initial CP degradation. Previous studies conducted in vivo have demonstrated both increased (19) and decreased (13) protein degradation due to prewilting. Heat production may bind protein to reduce the susceptibility to degradation (3), but the present study was conducted with mini silos in which heat was quickly dissipated. Hay was higher in initially degraded protein than freeze dried forages. This effect was probably due to degradation during wilting from plant enzyme activity (17). Bromegrass CP degradation was lower at 24 h for hay than for freeze dried forage. Previous work showed that in vivo protein degradation was reduced for dried compared to fresh forages (2, 28). Maturity x preservation method interaction effects were significant (P < .05) for TCA soluble protein before enzymatic degradation of alfalfa, bromegrass and com. This interaction was significant for alfalfa and canarygrass after 2 h degradation, but was only significant for alfalfa after 24 h degradation. Silage protein degradation fi— — ' ‘35» a Iwm‘ _1 54". "It'd: . %— _. 163 consistently declined with increased maturity of alfalfa, but hay CP degradation was higher in mid-maturity alfalfa than in early or late (Figure 1). The mid maturity alfalfa was harvested in a wet period, and was dried indoors at a slower rate than other forages. Forages increase uptake of nitrogen after rainfall (27), and proteolysis may have had greater opportunity to proceed during the slower drying. Maturity did not affect TCA soluble CP in freeze dried bromegrass that was not incubated with enzyme, though it affected other bromegrass samples (Figure 2). Freeze dried mid maturity canarygrass was higher in CP degradation at 2 h compared to early and late‘canarygrass. This contradicted other preserved canarygrass which tended to decline in degraded protein with increased maturity (Figure 3). Early freeze dried corn was lower than late freeze dried corn in TCA soluble protein before degradation with enzyme, but ensiled corn did not differ (Figure 4). 164 REFERENCES . Agricultural Research Council. 1984. Report of the Protein Group of the Agricultural Research Council Working Party on the Nutrient Requirements of Ruminants. Commonw. Agric. Bur., Slough, Engl. . Beever, D. B., S. B. Cammell, and A. Wallace. 1974. The digestion of fresh, frozen and dried perennial ryegrass. Proc. Nutr. Soc. 33: 73A. . Broderick, G. A., J. H. Yang, and R. G. Koegel. 1993. Effect of steam heating alfalfa hay on utilization by lactating dairy cows, J. Dairy Sci. 76: 165. . Brown, W. F. and W. D. Pitrnan. 1991. Concentration and degradation of nitrogen and fibre fractions in selected tr0pical grasses and legumes. Tropical Grasslands 25:305. . Filmer, D. G. 1982. Assessment of protein degradability of forage. Page 1 in Forage Protein in Ruminant Animal Prod. Occasional Publication No.6 Br. Soc. Anim. Prod. . Goering, H. K. and P. J. Van Soest. 1970. Forage Fiber Analyses (Apparatus, Reagents, Procedures, and Some Applications). Agric. Handbook No. 379. ARS- USDA, Washington, DC. . Greenhi11,W. L. 1959. The respiration drift of harvested pasture plants during drying. J. Sci. Food Agric. 10: 495. . Hach, C. H., S. V. Brayton, and A. B. Kopelove. 1985. A powerful Kjeldahl nitrogen method using peroxymonosulfuric acid. J. Agric. Food Chem. 33: 1117. . Hoffman, P. C., S. J. Sievert, R. D. Shaver, and D. K. Combs. 1992. In situ CP degradation kinetics of perennial forages. J. Dairy Sci. 75(Suppl. 1):210. 10. Janicki, F. J ., and C. C. Stallings. 1988 Degradation of crude protein in forages determined by in vitro and in situ procedures. J. Dairy Sci. 71:2440. 11. Kohn, R. A. and M. S. Allen. 1993. An in vitro method to measure protein degradation of feeds for ruminants using enzymes extracted from rumen contents. J. Dairy Sci. 76: Suppl. 12. Kwakkel, R. P. ,J. Van Bruchem, G. Hof, andH. Boer. 1986. The 1n sacco degradation of crude protein and cell wall constituents in grass, alfalfa and maize silages. Neth. J. Agric. Sci. 34:116. 13. Makoni, N. F., J. A. Shelford, and L. J. Fisher. 1989. Effect of wilting and maturity (3)8 1grass/legume silage on protein degradation in the rumen. J. Dairy Sci. 67 :Suppl. 165 14. McDonald, P. 1982. The effect of conservation processes on the nitrogenous 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. components of forages. Page 41 in Forage Protein in Ruminant Animal Prod. Occasional Publication No. 6 Br. Soc. of Anim. Prod. McKersie, B. D. 1981. Proteinases and peptidases of alfalfa herbage. Can. J. Plant Sci. 61: 53. Messman, M. A., W. P. Weiss, and D. O. Erickson. 1992. Effects of nitrogen fertilization and maturity of bromegrass on nitrogen and amino acid utilization by cows. J. Anim. Sci. 70: 566. Messman, M. A., W. P. Weiss, and M. E. Koch. 1992. Electrophoresis as a tool to ascertain effects of preservation method on forage protein composition. Page 35 in Ohio Agric. Experiment Station Research Highlights, Wooster, OH. Michalet-Doreau, B. and M. Y. Ould-bah. 1992. Influence of hay making on in situ nitrogen degradability of forages in cows. J. Dairy Sci. 75: 782. Narasimhalu, P., E. Teller, M. Vanbelle, M. Foulon, and F. Dasnoy. 1989. Apparent digestibility of nitrogen in rumen and whole tract of friesian cattle fed direct-cut and wilted grass silages. J. Dairy Sci. 72: 2055. National Research Council. 1989. Nutrient Requirements of Dairy Cattle. 6th rev. ed. Natl. Acad. Sci., Washington, DC. Nocek, J. E. 1988. In situ and other methods to estimate ruminal protein and energy digestibility: a review. J. Dairy Sci. 71: 2051. Petit, H. V. and G. F. Tremblay. 1992. In situ degradability of fresh grass and grass conserved under different harvesting methods. J. Dairy Sci. 75: 774. Rees, D.V.H. 1982. A discussion of sources of dry matter loss during the process of haymaking. J. Agric. Eng. Res. 27: 469. Robertson, J. B. and P. J. Van Soest. 1977. Dietary fibre estimation in concentrate feedstuffs. J. Anim. Sci. 45: Suppl. 254. Ruxton, I. B. and P. McDonald. 1974. The influence of oxygen on ensilage. 1. Laboratory studies. J. Sci. Food Agric. 12: 706. SAS Institute Inc. 1991. Statistical Visualization for the Macintosh. JMP ver. 2.2. SAS Institute Inc., Cary, NC. Thom, E. R., G. w. Sheath, and A. M. Bryant. 1989. Seasonal variations in total nonstructural carbohydrate and major element levels in perennial ryegrass and paspalum in a mixed pasture. New Zealand J. Agric. Res. 32: 157. Thomson, D. J ., E. Pfeffer, and D. G. Armstrong. 1969. Effects of drying grass on srtes of its digestion in sheep. Proc. Nutr. Soc. 28: 26A. 166 29. Vik-Mo, L. 1989. Degradability of forages in sacco 2. Silages of grasses and red clover at two cutting times, with formic acid and without additive. Acta Agric. Scand. 39: 53. 30. Voss, V. L., D. Stehr, L. D. Satter, and G. A. Broderick. 1988. Feeding lactating dairy cows proteins resistant to ruminal degradation. J. Dairy Sci. 71: 2428. 167 TABLE 1. Mean forage composition by species and maturity level across forage preservation methods (11 = 12). Item CP NDF ADF Lignin Ash % Alfalfa Early 27.0 35.7 25.9 5.1 9.6 Mid 20.1 48.7 35.2 8.0 8.5 Late 16.5 52.6 38.1 9.0 9.0 SEM .22 .48 .42 .11 . 10 Bromegrass Early 19.4 53.7 29.0 3.1 8.6 Mid 14.3 63.7 35.9 4.7 7.6 Late 12.3 66.1 37.7 5.6 7.7 SEM .17 .27 . 19 .06 .07 Canarygrass Early 20.8 53.9 28.5 2.7 8.1 Mid. 16.6 . 64.6 34.3 4.1 7.8 Late 13.9 65.5 35.5 4.8 7.4 SEM .22 .41 .28 .09 .14 Corn ’ Early 7.3 46.6 25.5 2.7 3.7 Late. 7.4 42.3 22.3 2.5 3.4 SEM .13 .74 .47 .08 .08 lSequential extraction was used for determination of NDF, ADF and sulfuric acid lignin. 168 TABLE 2. Mean protein degradation at 0, 2 and 24 h by forage species and maturity level across forage preservation methods.1 Forage 0 h 2 h 24 h % Alfalfa Early 39.3 55.4 63.7 Mid 36.6 49.8 57.5 Late 30.8 42.6 51.6 P, early vs. mid?- .05 .0001 .0001 P, mid vs. late .0001 .0001 .0001 Bromegrass Early 25.6 39.9 47.8 Mid 21.3 31.6 39.7 Late 17.6 25.3 36.5 P, early vs. mid .0001 .0001 .0001 P, mid vs. late .0001 .0003 .05 Canarygrass Early 28.6 39.5 48.3 Mid 27.3 39.0 47.0 Late 28.6 37.5 45.3 P, early vs. mid NS NS NS P, mid vs. late NS NS NS Corn Early 25.4 26.6 36.2 Late 24.7 26.6 34.9 P, early vs. late NS NS NS 1Values expressed as percentage CP soluble in TCA after incubation with ruminal enzymes for time indicated. For perennials n = 12, com 11 = 9. 2Significance of contrast, NS indicates P > .1. 169 TABLE 3. Mean protein degradation at 0, 2 and 24 h by forage species and preservation method across maturities.1 Forage 0 h 2 h 24 h % Alfalfa Freeze dried 19.8 39.1 52.7 Hay - 24.2 42.6 56.5 Silage (30% DM) 49.2 57.3 59.9 Silage (38% DM) 49.1 58.0 61.3 P, dried vs. ensiled2 .0001 .0001 .0001 P, freeze dried vs. hay .01 .002 .001 P, wet vs. dry silage NS NS NS Bromegrass Freeze dried 10.0 25.6 39.1 Hay 15.0 27.7 34.8 Silage (33% DM) 29.7 36.5 45.2 Silage (44% DM) 31.4 39.3 46.3 P, dried vs. ensiled .0001 , .0001 .0001 P, freeze dried vs. hay .0001 NS .03 P, wet vs. dry silage .01 NS NS Canarygrass Freeze dried 12.8 27.6 40.4 Hay 20.9 34.2 41.9 Silage (34% DM) 42.8 47.3 53.6 Silage (43% DM) 36.1 45.5 51.5 P, dried vs. ensiled .0001 .0001 .0001 P, freeze dried vs. hay _ .01 .0001 NS P, wet vs. dry silage .05 NS NS Corn Freeze dried 11.8 21.7 29.0 Silage (31% DM) 32.5 30.8 39.4 Silage (38% DM) 30.8 27.4 38.3 P, dried vs. ensiled .0001 .001 .002 .05 NS NS P, wet vs. dry silage 1Values expressed as CP soluble in TCA after incubation with ruminal enzymes for time indicated. For perennials n = 9; corn n = 6. 2Significance of contrast, NS indicates P > .1. 170 Figure 1. Percentage of alfalfa forage protein degraded in early mid and late maturity forage. I Indicates freeze dried sample, I hay, 5 low DM silage, and D high DM silage. A: Percentage soluble at 0 h, B: 2 h and C: 24 h. ) 171 (A 1'11 Early m w a a L L M m M m M y y H H a a E E 4.. 0 5 0. 4.. 0 5 0 4.. 0. 5 0 5 0 4.. 0 .... 0 7 6 4 3 1 7 6 4 3 1 7 6 4 3 1 Aoev 5395 25.5 ecumewon 36v 5889 25.5 coeaLweQ 36v £32m 32.5 33..on 172 Figure 2. Percentage of bromegrass forage protein degraded in early mid and late maturity forage. I Indicates freeze dried sample, 3 hay, E3 low DM silage, and U high DM silage. A: Percentage soluble at 0 h, B: 2 h and C: 24 h. 173 ) A ( 5 0 5 0 5 0 7 6 4 3 1 $3 589:. 35.5 @223on a, and Late Mid Early Late Mid Early 5 0 5 7 6 4 30 15- 0 Aoev 5395 25.5 mega—wen— ) C ( 0 5 0 3 1 45- ?ev 5395 35.5 movemen— Late Mid Early 174 Figure 3. Percentage of reed canarygrass forage protein degraded in early mid and late maturity forage. I Indicates freeze dried sample, . hay, '3 low DM silage, and U high DM silage. A: Percentage soluble at 0 h, B: 2 h and C: 24 h. 175 A) ( Late Late Mid B) ( Early r 5 0 5 50 050 5 1 31 7 6 4 5 0 5 0 7 6 4 3 $3 £39:— 339 Boe.—us: $3 589:. 25.5 02:29: 33 589:. 25.6 evasion 176 Figure 4. Percentage of whole plant corn protein degraded in early and late maturity forage. I Indicates freeze dried sample, 1!?! low DM silage, and 0 high DM silage. A: Percentage soluble at 0 h, B: 2 h and C: 24 h. Degraded crude proteln (%) Degraded crude protein (95) Degraded crude protein (%) 178 CHAPTER 7: PREDICTION OF PROTEIN DEGRADATION OF FORAGES FROM SOLUBILITY FRACTIONS ABSTRACT Two experiments were conducted to determine the relationship between ruminal digestion of forage protein and fractionation based on solubility. The first experiment used 42 forages (each replicated 3 times) including different species, maturities, and means of conservation. Crude protein was fractionated into 6 parts for each forage: trichloroacetic acid soluble, buffer soluble, acetone soluble, neutral detergent soluble, acid detergent soluble and insoluble. Crude protein degradation at 2 and 24 h (determined previously) was regressed against solubility fractions. Protein fractionation patterns differed among forages due to effects of species, maturity and means of conservation. Multiple regression analysis resulted in prediction of crude protein degradation at 2 and 24 h with R2 of .88 and .81 respectively. Higher levels of buffer and neutral detergent soluble fractions were associated with higher protein degradation, while higher levels of acetone soluble, acid detergent soluble, and insoluble crude protein were associated with lower degradation. In the second experiment, eight forages (each replicated twice) were digested with ruminal enzyme for 0, 2, 6 and 24 h, and then were extracted as described. Solubility in TCA and acetone increased during degradation, while buffer soluble, acid detergent soluble and insoluble CP decreased during degradation. In both experiments, buffer soluble CP was the only uniform fraction, while other fractions contained proteins that degraded at diverse rates. Solubility of CP is related to degradation properties, but further research 179 is needed to isolate uniform fractions for use in models that predict CP degradation rates. (Key words: protein degradation, protein solubility, forages) Abbreviation key: TCA = Trichloroacetic acid. INTRODUCTION The rate and degree of ruminal protein degradation of feeds is an important consideration for formulating dairy cattle rations (22). Forages provide much of the protein consumed by dairy cattle, and the degradation of forage protein varies depending on plant species, maturity, conservation method (13) and season (1). Since forages are an important variant, a routine method of analysis or prediction of forage protein degradation is required to balance dairy cattle rations accurately. Rumen degradation of forage protein depends on rate of degradation of individual proteins in the rumen, and on rate of passage of those proteins from the rumen. Both of these factors may be influenced by protein associations with other feed components, and by availability of chemical bonds of the proteins to enzymatic attack. Separation of feed proteins based on solubility may have practical significance to ruminal protein degradation, since ultimately, the solubility of a protein determines the microenvironment, or feed componerits, with which it associates (i.e. aqueous, lipid, etc.), and this association may alter the availability of peptide bonds, or alter the passage rate of the protein. Additionally, solubility may relate to other aspects of protein structure which affect protein degradation. Various researchers have attempted to quantify ruminally degraded protein based on solubility properties in saline solutions, buffers or detergents. Buffer solutions dissolve most non-protein nitrogen and varying amounts of true protein (16). In general, buffer-soluble true protein has been shown to be more rapidly degraded by bacterial protease and by ruminal microorganisms than insoluble protein, but the rate of 180 degradation varies among protein sources and some soluble proteins are not rapidly degraded_(l7, 23). Insoluble forage crude protein includes hydrophobic compounds such as chlorophyll that contain non-protein nitrogen, and are selectively extracted in acetone (31). In addition, membranes contain amphiphilic true proteins that may be protected from degradation by the lipid bilayer (16). These proteins can be extracted in detergent solutions. Detergent solution at pH 7 extracts cell contents and leaves CP associated with plant cell wall (28), which is degraded slowly in the rumen (15). Acid detergent insoluble crude protein is a measure of heat damage that renders protein indigestible (32, 29). The objectives of the present study were: 1) to compare changes in protein solubility fractions across forage species, maturities and preservation methods, 2) to determine rates of degradation of individual protein fractions, and 3) to use this information to evaluate the possibilities for predicting forage protein degradation from solubility fractions. METHODS Forages and Protein Degradation Forages used in this study were described elsewhere (13). These forages included alfalfa (Medicago sativa L), smooth bromegrass (Bromus inerrnis Leyss), reed canarygrass (Phalaris arundinacea L), and whole plant corn (Zea mays L) at different stages of maturity. Samples of each forage were freeze dried immediately after harvesting, and others were preserved as hay and silage at high and low DM content. Three field replicates were taken for each of 42 forage types. Protein degradation was previously determined for each replicate forage by incubation with a crude extract of ruminal enzyme for 2 and 24 h (13). Enzymes were extracted from rumen fluid by solubilization in water after releasing them from 181 microbes by extraction in butanol and acetone. After incubation with rumen fluid extract, undegraded protein was determined as that which precipitated in trichloroacetic acid (I CA) at a final concentration of 80 g/L (12). Protein Fractionation Each forage replicate was subjected to sequential extraction of crude protein according to the scheme depicted in Figure 1. Forage samples containing 16 mg of N were weighed into 50 ml centrifuge tubes, which were then randomized before extraction. Tubes were placed under C02 and 30 ml reduced media (12) was added. The media was a pH 6.8 bicarbonate buffer with minerals and reducing agent ( NazS.9H20). Tubes were purged with C02 while adding media, and then were capped, shaken and stored for 2 h at 4 °C. Reduced media was used to prevent precipitation of forage protein due to enzymatic reactions that were demonstrated previously in bromegrass (14). Preliminary research (11) indicated that storage at 4°C resulted in solubilization of crude protein without significant degradation of that protein by plant enzymes. Storage at low temperature resulted in reduced NPN formation over time compared to storage at 38°C, and was comparable to samples stored with enzyme inhibitor (phenymethylsulfonyl fluoride). After 2 h in buffered media, samples were centrifuged at 4°C and 1500 g, for 20 minutes. The supernatant was removed to different centrifuge tubes, 15 ml of 240 g/L TCA was added, and these tubes were centlifuged at 1500 g, for 20 minutes. The supernatant from this extraction was presumed to be soluble NPN (ammonia, nitrate, amino acids, peptides) or "fraction 1", and pellet was buffer soluble true protein (fraction 2) (18). The first pellet containing most of the original forage was extracted with 80% acetone (20% water, vol/vol) for 2 h. Acetone (40 ml) was added and the samples shaken vigorously three times while stored at 25°C. These samples were centrifuged without chilling at 1500 x g for 20 min. The supernatant was removed to different 182 centrifuge tubes and the acetone evaporated over N2 in a sand bath at 40°C. This fraction was presumed to be acetone soluble NPN (fraction 3) which is composed of strongly hydrophobic pigments and membrane-associated compounds (i.e. chlorophyll) (7 ). The residue was next extracted with pH-7 detergent. The detergent solution was adapted from Goering and Van Soest (9) with modification of the solution by substitution of triethylene glycol for 2-ethoxyethanol (18 g/kg) and the omission of decahydronaphthalene and sodium sulfite (25). Additionally EDTA was replaced with sodium oxalate (25 g/L) in order to remove nitrogen from the detergent solution, yet still provide a calcium chelator to solublize pectins. This detergent solution (40 ml) was added to the residual forage samples and incubated for l h at 95°C. It was cooled to 50°C and centrifuged at 1500 x g for 20 min. The supernatant was removed and presumed to be cell-wall associated crude protein (fraction 4). The pellet was extracted again using a similar method except for replacement of pH-7 detergent with sulfuric acid and detergent solution. Triton-X 100 (2 mlIL, Sigma Chemical Co., Saint Louis, MO) was added to 1 N sulfuric acid. This solution (40 ml) was added to the residue of forage samples, which were incubated for 1 h at 95°C before centrifugation. When the detergent cooled before and during centrifugation, a film formed at the top of the supernatant and adhered to the sides of the tube. This film was included with the supernatant fraction. The supernatant was presumed to be cell- wall-associated crude protein (fraction 5) and the pellet was presumed to be bound protein (fraction 6) (32, 29). Each of these protein fractions was subjected to a modified Kjeldahl analysis for nitrogen (10). Crude protein in each fraction was determined as N x 6.25. This value was expressed as a fraction of the total crude protein for the replicate (determined 183 previously). The percent yield was determined as 100 times the total of all six fractions. Statistics for the First Experiment Statistics were performed using JMP (26). For comparisons of forages for solubility fractions, each forage species was analyzed separately due to significant interaction effects of species with maturity and preservation method. Maturity, preservation method and the interaction effects were determined using AN OVA at each length of incubation. Further elucidation of main effects was made using the following Specific contrasts: early vs. mid maturity, mid vs. late maturity, freeze dried and field dried vs. ensiled forages, freeze dried vs. field dried, and wet vs. dry silage. The portion of each protein fraction that was degraded at 2 and 24 h was estimated from results of multiple regression. Several models were compared to determine rates of degradation for protein fractions that were theoretically possible. The model determined to be most appropriate a priori was as follows: 6 Y = 2 fi (X 3 i=1 . The portion of protein degraded (I’CA soluble) at 2 or 24 h is represented by Y, Xi represents the pool size of each solubility fraction (i) expressed as a fraction of total crude protein (determined previously), and fi represents the fractional portion of cnrde protein that is degraded at 2 or 24 h for each solubility fraction (1 to 6). The values of fi were determined as the slopes from multiple regression of degraded protein at 2 or 24 h by the initial solubility fractions. The intercept was fixed at zero for the regression because this is theoretically required, and allowing it to vary altered the prediction of intercepts. Adding this fixed value inflates the value of R2 (because it is adding a point at the low end of the curve), therefore R2 was determined separately from plots of observed vs. predicted degradation when using the previously calculated slopes. ‘l 184 Alternative models fixed negative slopes at 0 to determine other slopes more accurately, or fractions were pooled where separation of fractions was not believed to be adequate. Rate of protein degradation (k) was determined on the undegraded residue (undg = 1 - degraded) as follows: k = [1n (undg at time t + At) - 1n (undg at time t )] IAt, where t + At represents the duration of degradation (2 or 24 h) and t represents the previous determination (0 or 2 h), and At is the time interval involved in the measurement (2 or 22 h). Solubility Fractions After Degradation Eight forages were subjected to a second experiment where crude protein was fractionated after degradation in enzyme for 0, 2, 6, or 24 h. Forages used were a subset of those in the previous experiment: early alfalfa that was freeze dried, mid- maturity alfalfa hay, late alfalfa silage, early bromegrass silage, late bromegrass hay, mid-maturity canarygrass that was freeze dried, mid-maturity canarygrass silage, and late corn silage. Forages were incubated with ruminal enzyme as described previously, and after incubation with enzyme crude protein was fractionated using a modified procedure to that described previously. Since extraction required twice the amount of sample as was used in the degradation method, two samples of each replicate were incubated in different centrifuge tubes, and these were combined during the extraction. At the end of the incubation with enzyme, samples were cooled to 4°C and centrifuged at 1500 x g for 20 min. The supernatants were decanted to clean tubes and the pellets were combined for subsequent extractions carried out as described previously. Buffer soluble true protein was precipitated from the supernatant by addition of TCA. Nitrogen was determined on combined pellets. Crude protein soluble in TCA was determined indirectly by subtracting the total of all other fractions from the total CP which was determined previously. This was done because enzyme was TCA soluble so that this fraction could not be determined directly. All analyses were determined in 6 185 runs with the same rumen extract (2 replicates of 3 lengths of incubation were determined separately). Percentage of each protein fraction that was degraded was determined as 100 times the crude protein in the fraction at a point in time divided by the initial crude protein in that fraction. Effects of forage type and length of incubation were evaluated by analysis of variance for protein degraded within each fraction. Since interaction effects were significant, forages were compared for percentage protein degraded in each fraction at each length of incubation . Dried were compared to ensiled forages by student t test. Noting differences, similar comparisons were made among ensiled and dried forages. The Lucas test of nutritional uniformity was applied to estimate percentage protein degraded in each fraction and variation across forages (30). This procedure is based on the principle that the apparent degradation of a nutrient (Dax) is a function of the true degradation of the nutrient (Dtx), metabolic contribution of the nutrient (Mx), and amount of the nutrient in the diet (Xi) as follows: Dax = Dtx - Mx I Xi. Multiplying through by Xi gives the equation for a line: Dax(Xi) = Dtx(Xi) - Mx. If Dax(Xi) is plotted on the abscissa and Xi on the origin, the slope of the line is the true degradation and the negative intercept is the metabolic contribution of the nutrient. The R2 of the line indicates the uniformity of degradation of the fraction across diets. In the present experiment, amount of each fraction in the forage sample was plotted on the origin and this times the percentage degradation of the fraction was plotted on the abscissa. Separate plots were drawn for each length of incubation RESULTS AND DISCUSSION Fractionation of Forages Results of protein fractionation are presented in Tables 1 and 2. Most of the nitrogen (N) was soluble in TCA, neutral detergent or was insoluble. Corn also contained a 186 large amount of acetone soluble N (Table 1). While the acetone soluble fraction was dark green for other forages, it was bright yellow for corn. Corn proteins (especially zein) may degrade to hydrophobic peptides that resist ruminal degradation. It was presumed to be composed of chlorophyll or other pigments and may be indigestible NPN because compounds would have to be highly non-polar to dissolve in acetone. Such compounds are likely to contain long hydrocarbon chains but not oxygen. It would therefore be difficult for bacteria to use these compounds in the anaerobic environment of the rumen. Ensiling of feeds resulted in several important changes in protein fractionation (Table 2). Crude protein soluble in TCA increased for all forages due to ensiling. This result would have occurred due to the degradation of forage protein to peptides, amino acids or ammonia which is expected during ensiling (19). Acetone soluble N also increased for silages compared to dried forages. Several speculations are plausible to explain this effect. It may have resulted from an inadequate extraction of pigments in the dried forages, but ensiling may release these pigments from the plant structure. Alternatively, this fraction may contain hydrophobic peptides which were insoluble in buffer and therefore not extracted by TCA. Buffer soluble true protein was greatly reduced in ensiled compared to dried forages. Previous work has demonstrated the rapid degradation of soluble protein in alfalfa (6). Neutral detergent soluble protein (fraction 4) was also reduced considerably by ensiling, and acid detergent soluble (fraction 5) and insoluble (fraction 6) protein were reduced less dramatically. The fact that nearly all protein fractions were reduced in ensiled compared to dried forages demonstrates that even buffer insoluble CP is susceptible to protein degradation during ensiling. 187 Prediction of Degradation from Fractions The prediction of protein degradation at 2 h from multiple regression analysis is shown in Table 3. Including all fractions in the model resulted in an R2 of .88 (model 1). The intercept was highly significant. perhaps due to multiple pools within some fractions. If the size of each fraction (xi) were 0, one would expect the degraded protein at 2 or 24 h to also be 0. This was the grounds for forcing the intercept through 0 for model 2. The objective was to estimate the percentage protein degraded for each fraction, not necessarily provide for the best fit line. The slopes presented in model 2 should be better estimates of percentage protein degraded at 2 h than those from model 1 because the intercept is not allowed to bias the prediction. About 100% of buffer soluble N (fractions 1 and 2) is predicted to be degraded at 2 h. Since TCA solubility is the measurement of degradation in these studies this estimate is expected. The fact that 100% of buffer soluble protein is degraded is consistent with results comparing ensiled and dried forages. Fraction 4 was the next most degraded at 2 or 24 h, again confirming results from comparison of ensiled and dried forages. Acid detergent soluble and insoluble crude protein were not predicted to be degraded at 2 h. The estimate of acetone soluble crude protein degradation was negative. If the difference in amount of acetone soluble crude protein resulted from differing efficiency of extraction on different feed types, as has been postulated, then the higher extraction in acetone would be associated with a lower extraction of another fraction. It is likely that pigments not removed by acetone would be removed by detergent solution, so an underestimation of fraction 3 could result in an overestimation of fraction 4. Since fraction 4 is degraded, and 3 may not be, the overestimation of degradation due to high levels of fraction 4 would be balanced by a negative slope associated with fraction 3. Further experimentation with the model was based on three considerations. First, most of the crude protein was found in fractions 1, 4 and 6. Second, the estimates 188 for the smaller fractions 3 and 5 (Tables 3 and 4) were negative, though we can not expect negative degradation of crude protein. Third, coefficients of variation were high for fractions 3 and 5. Models 3 and 4 were designed to minimize the contribution from fractions 3 and 5 to the slopes of the other fractions by either pooling fractions (model 3) or forcing the slopes of fractions 3 and 5 to 0 (model 4). The results are consistent with previous results. Fractions 1 and 2 again were nearly completely degraded at 2 h. Fractions 3 and 4 together (model 3) were predicted to degrade more slowly than fraction 4 by itself (model 4). This demonstrates the importance of removing fraction 3 from fraction 4, and if this were done more thoroughly, the degradation of fraction 4 may have been higher. Crude protein degradation at 24 h was estimated for each fraction using the models described for degradation at 2 h (Table 4). The strong relationship between solubility and degradation at 24 h (Figure 2) indicates the potential for analyzing forages routinely based on solubility. These results are consistent with those for degradation at 2 h. Model 4 estimated fractions 1 and 2 to be nearly completely degraded, fraction 4 to be 60% degraded, fraction 6 to be 11% degraded, with fractions 3 and 5 set to 0% degradation. The rate of degradation of fraction 4 is estimated from the percentage degraded at 2 h (I' able 3, model 4) as 13.5%Ih, while it is estimated as 3.8 %Ih over the course of 24 h (Table 4, model 4). Likewise, the rate of degradation of fraction 6 is estimated as 1.7%Ih vs. .5%Ih over 2 and 24 h respectively. The estimates for rate of degradation declines from fast to slow over time. This change in degradation rate indicates the probability of multiple fractions within fraction 4 and possibly 6 which degrade at different rates. Further work may be warranted to investigate further separation, perhaps exploiting better removal of acetone soluble fraction. 189 Solubility Fractions After Degradation Change in protein fractions due to degradation with ruminal enzyme is shown (Figure 3). Crude protein soluble in TCA and acetone increased over time in dried forages. Though acetone soluble crude protein also increased over time for ensiled forages, TCA soluble crude protein actually decreased. Amount of neutral detergent soluble crude protein did not change due to incubation with enzyme. All other fractions decreased when incubated with enzyme. In the first experiment, TCA soluble crude protein increased due to protein degradation of all forages including silage. The difference in the present experiment may relate to the measurement of TCA soluble protein. When crude protein degradation was measured in forages in the previous experiment, TCA was added directly to in vitro media after incubation with ruminal enzyme. In the present experiment, TCA was added to the supernatant from the degraded forage only after centrifugation of the residue from degradation. Thus crude protein that was insoluble in media was not available to be solubilized in TCA. On the other hand, if TCA is added directly to feed proteins, especially when hydrophobic crude protein (pigments or peptides) have been released due to digestion, the TCA may solubilize some of these compounds. However, if TCA is added only to the buffer soluble crude protein, the strongly hydrophobic compounds have already been removed, and therefore are not candidates for TCA solubility. Hydmphobic compounds would be solubilized in acetone. This solution was chosen to solubilize pigments and other NPN compounds that would otherwise be contained in the detergent soluble fraction. Observation that the neutral detergent soluble fraction does not decrease due to digestion, while acetone soluble CP increases, suggests that this CP may be extracted from another fraction. It may contain hydrophobic or glycated peptides. In fact it may relate to the soluble protein identified 190 in other experiments (apparently peptides) that appears to accumulate due to ruminal digestion (5). A better understanding of acetone soluble crude protein is needed to evaluate whether or not it should be included in the degraded, undegraded, or indigestible fraction, especially considering the fact that after degradation it comprises 15 to 20 % of the total crude protein (Figure 3). Buffer soluble crude protein (fractions 1 and 2) was nearly completely degraded by 2 h (Figure 3, Table 5). This observation is consistent with the discussion of the previous experiment. However, the fact that fraction 4 was apparently not affected by degradation contradicts the changes from dried to ensiled feeds, and the predictions from regression described in the previous experiment. A possible explanation may be that though protein from this fraction was degraded, crude protein from other fractions may have been made soluble in neutral detergent by degradation, thus negating the previous effect. This protein may have been supplied from the acid detergent soluble or insoluble fractions, which did show disappearance in this experiment. Though protein fractions appear to be degraded by enzyme at the initial stages, this degradation is slowed or stopped at the later time points (Figure 3 and Table 6). Buffer soluble protein is the only fraction that was nearly completely degraded. The enzymeused in these experiments appears to maintain its activity on azocascin for extended incubations, and addition of fresh enzyme after 8 h incubation did not increase soybean meal protein degradation rate (13). Therefore, the slowing rate of degradation is likely to be a function of the forage and not only the method of analysis. Plant structures associated with proteins may inhibit degradation of some forage protein fractions using this system. Forages varied in the percentage protein degraded at each length of incubation for most fractions. These effects are shown for 2 h degradation (Table 5). Much of the variation was between dried forages and silage. Since much of 191 the readily degraded protein was already lost from the silage samples, the remaining fraction would degrade at a slower rate. The Lucas test is an efficient means to estimate the percentage protein degraded within a fraction across forages, and the variation associated with that estimate (30). Fraction 2 protein was shown to be entirely degraded within 2 h across all forages (Figure 4A, Table 7). The small but significant negative intercept probably was due to contamination of protein from other fractions, or to a constant amount of undegraded crude protein associated with this fraction for all forages. During the extraction, insoluble suspensions were observed after centrifugation for buffer soluble protein. These suspensions were precipitated by addition of TCA. Most likely these particles caused this negative intercept. Other fractions appeared less uniform for degradation of protein when analyzed using the Lucas test. The percentage degradation was significant at 24 h for fraction 4 (Figure 4B, Table 7 ). None-the-less there was considerable variation across forages and a low percentage degradation. Degradation of acid detergent soluble and insoluble fractions appeared to be greater than that of neutral detergent soluble, and were more consistent across forages (Figure 5, Table 7). The accuracy of these estimates of percentage degradation is questionable where the intercept is negative. When the Lucas test is applied to animal digestibility data, the negative intercept refers to metabolic protein. In the present system, this CP could be contributed from other fractions. However, it is more likely to be due to non-uniform fractions. The most important differences in protein degradation and in amount of CP in a fraction are due to effects of ensiling. The ensiled forages have lower amounts of some fractions, and degradation of those fractions is slower because the more readily degraded portion was lost during ensiling. Therefore, with certain fractions, lower amounts of the fraction may be —1—_—v—Tn_—r '-~ ; ' .2: .q._..=:-.:>_.-_ —.;:«.r§:_-“_~j “... 192 associated with slower degradation. This would result in an artificially lower intercept, and an inflated slope for the line (Figure 5). Comparison to Previous Results Previous efforts to predict protein degradation of forages from solubility have provided mixed results. Buffer soluble protein has been shown to be rapidly degraded by proteolytic enzymes (6), but since it comprises a small amount of the total CP, a relatively high amount of buffer soluble protein does not necessarily prescribe a high degradability of the forage protein (24). Likewise, neutral detergent insoluble protein was shown to degrade at a slow rate for some forages degraded in situ (15), but this fraction also represented a small portion of the total CP. Acid detergent insoluble crude protein is used as an indicator of heat damage that renders the protein indigestible (8, 29), but some studies have shown a degree of degradation of this fraction (4, 33). - Other researchers have investigated the possibility of using gel electrophoresis to isolate forage proteins for the prediction of protein degradation (20). This work demonstrated that even though the exact same protein may be isolated from different forages (i.e. ribulose-bis-phosphate-carboxylase/oxygenase), its rate of degradation may vary from forage to forage (20), perhaps due to protein associations with other forage components. For example, tannins have been implicated in the slowing of degradation of forage protein (3), and cellulase treatment increased degradation of forage protein in vitro (13). Therefore, measuring the amount of a specific protein fraction may not be as important to predicting overall degradation as measuring the degree of association of forage proteins with tannins, fiber, or membranes. The methods used to fractionate protein and determine degradation may affect the results. In the present study, forages were freeze dried and resolubilized in reduced media. This drying process was chosen because it was determined to decrease changes in solubility and degradation compared to freezing and thawing [Chapter 2], (14) or 193 oven drying (2), however, results may have differed if fresh feeds were analyzed. The present study also utilized a sequential extraction so that independent fractions would be obtained. Nitrogen was removed from detergent solutions to make sequential extraction possible. Solubility was determined by centrifugation ratherthan filtration, thus potentially decreasing the overall solubility. The method of measuring protein degradation also differed from some previous work in that TCA insoluble protein was regarded as the undegraded protein and microbial contamination was eliminated by using extracted enzyme. Other methods may over-estimate the undegraded crude protein due to microbial attachment to feeds (21), or underestimate it because protein is considered degraded as soon as it is reduced in particle size enough to leave the nylon bag in the rumen, or filter through filter paper used for in vitro methods (27). On the other hand, the method used in the present study may be limited by the enzyme profile that was extracted and the possible importance of live microbes to the degradation. Implications The’actual amount of degraded and undegraded protein supplied by a forage depends on both degradation and passage rates. The protein fractions isolated in this study may in fact pass from the rumen at different rates owing to their solubility and associations with certain plant tissues. Further research could ultimately lead to the isolation-of protein fractions that degrade and pass at constant and uniform rates, and would enable accurate prediction of protein degradation in vivo. This is the first time all of the CP of forages was extracted sequentially to determine the possibility of predicting protein degradation from a complete array of solubility fractions. The prediction of protein degradation at 2 h or 24 h indicates a potential for use of solubility to predict degradation. Buffer soluble protein was degraded rapidly in all forages, and acetone soluble CP appeared not to be degraded. Other fractions degraded to differing extents depending on the forage, and much of this 194 inconsistency appears to have resulted from the presence of multiple pools within certain fractions. Prediction of protein degradation from solubility fractions can be improved by isolating more homogenous fractions with respect to protein composition and interactions of the proteins with other feed components. Further research is needed to fractionate proteins into pools that degrade completely at the same rate. 10. ll. 12. 13. 14. 15. 195 REFERENCES Abdalla,H. O., D. G. Fox, andR. R. Seaney. 1988. Variationinprotein and fiber fractions 1n pasture during the grazing season. J. Anim. Sci. 66: 2663. Abdalla, H. O., D. G. Fox, and P. J. Van Soest. 1988. An evaluation of methods for preserving fresh forage samples before protein fraction determinations. J. Anim. Sci. 66: 2646. Barry, T. N. and T. W. Manley. 1984. The role of condensed tannins in the nutritional value of Lotus pendunculatus for sheep. Br. J. Nutr. 51: 493. Broderick, G. A., J. H. Yang, and R. G. Koegel. 1993. Effect of steam heating alfalfa hay on utilization by lactating dairy cows. J. Dairy Sci. 76. 165. Chen, G. H. J. Strobe], J. B. Russell, andC. J. Sniffen. 1987. Effect of hydrophobicity on utilization of peptides by ruminal bacteria 1n vitro. Appl. Environ. Microbiol. 53: 2021. Chemey, D. J. R., J. J. Volenec, and J. H. Cherney. 1992. Protein solubility and degradation in vitro as influenced by buffer and maturity of alfalfa. Anim. Feed Sci. Technol. 37:9. Dietz, K. J. 1989. Leaf and chloroplast development in relation to nutrient availability. J. Plant Physiol. 134: S44. Goering, H. K, C. H. Gordon, R. W. Hemken, D. R. Waldo, P. J. Van Soest, and L. W. Smith. 1972. Analytical estimates of nitrogen digestibility 1n heat damaged forages. J. Dairy Sci. 55: 1275. Goering, H. K. andP. J. Van Soest. 1970. Forage Fiber Analyses (Ap Mrspara Reagents, Procedures, and Some Applications). Agric. Handbook No. 379. ARS- USDA, Washington, DC. Hach, C. H., S. V. Brayton, and A. B. Kopelove. 1985. A powerful kjeldahl nitrogen method using peroxymonosulfuric acid. J. Agric. Food Chem. 33: 1117. Kohn, R. A. 1993. In vitro methods to determine forage protein degradation in the rumen. Ph. D. Diss., Michigan State Univ., East Lansing. Kohn, R. A. and M. S. Allen. An 1n vitro method to measure protein degradation of feeds using enzymes extracted from rumen contents. Br. J. Nutr. Submitted. Kohn, R. A. and M. S. Allen. Effect of plant maturity and preservation method on in vitro protein degradation of forages. J. Dairy Sci. submitted. Kohn, R. A. and M. S. Allen. 1992. Storage of fresh and ensiled forages by freezing affects fibre and crude protein fractions. J. Sci. Food Agric. 58: 215. Lindberg, J. E. 1988. Influence of cutting time and N fertilization of the nutritive value of timothy. 2. Estimates of rumen degradability of nitrogenous compounds" Swedish J. Agric. Res. 18:85. 16. 17 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 196 Lyttleton, J. W. 1973. Proteins and nucleic acids. Pages 63-103 in Chemistry and Biochemistry of Herbage. ed. G. W. Butter and R. W. Baily. Academic Press, New York, NY. Mahadevan, S., J. D. Erfle, and F. D. Sauer. 1980. Degradation of soluble and insoluble proteins by Bacteroides amylophilus protease and by rumen microorganisms. J. Anim. Sci. 50:723. Marals, J. P. and T. K. Evenwell. 1983. The use of trichloroacetic acid as precipitant for the determination of "true protein" in animal feeds. South African J. of Anim. Sci. 13: 138. McKersie, B. D. 1981. Proteinases and peptidases of alfalfa herbage. Can. J. Plant Sci. 61: 53. Messman, M. A., W. P. Weiss, and M. E. Koch. 1992. Electrophoresis as a tool to ascertain effects of preservation method on forage protein composition. Pages 35 to 39 in Ohio Agric. Experiment Station Research Highlights, Wooster, OH. Michalet—Doreau, B. and M. Y. Ouldbah. 1992. In vitro and in sacco methods for the estimation of dietary nitrogen degradability in the rumen - a review. Anim. Feed Sci. Technol. 40: 57. National Research Council. 1989. Nutrient Requirements of Dairy Cattle. 6th rev. ed. Natl. Acad. Sci., Washington, DC. Nocek, J. B., J. H. Herbein, and C. E. Polan. 1983. Total amino acid release rates of soluble and insoluble protein fractions and concentrate feedstuffs by Streptomyces griseus. J. Dairy Sci. 66: 1663. Pion, R., C. Genest, G. Bayle, and P. Thivend. 1983. Assessment of protein degradability in concentrates using an enzymatic method. Pages 207-210 in IV'th Int. Symp. Protein Metabolism and Nutrition, Clermont-Ferrand (France), Sept. 5— 9, 1983. Inst. Natl. Rech Agron. Publ. Robertson, J. B. and P. J. Van Soest. 1977. Dietary fibre estimation in concentrate feedstuffs. J. Anim. Sci. 45: Suppl.254 SAS Institute Inc. 1991. Statistical Visualization for the Macintosh. SAS Institute Inc., Cary, NC. Spencer, D., T. J. V. Higgins, M. Freer, H. Dove, and J. B. Coombe. 1988. Monitoring the fate of dietary proteins in rumen fluid using gel electrophoresis. Br. J. Nutr. 60: 241. Van Soest, P. J. 1964. Symposium on nutrition and forage and pastures: new chemical procedures for evaluating forages. J. Anim. Sci. 23: 838. Van Soest, P. J. 1965. Use of detergents in analysis of fibrous feeds. 111. Study of effects of heating and drying on yield of fiber and lignin in forages. J. Assoc. Offic. Anal. Chem. 48:785. ‘l 30. 31. 32. 33. 197 Van Soest, P. J. 1982. Chapter 4. Fecal composition, mathematics of digestion balances and markers. Pages 39-57 in Nutritional Ecology of the Ruminant, Cornell University Press, Ithaca, NY. Waghom, G. C., I. D. Shelton, and V. J. Thomas. 1989. Particle breakdown and rumen digestion of fresh ryegrass (Lolium perenne L.) and luceme (Medicago- sativa L.) fed to cows during a restricted feeding period. Br. J. Nutr. 61: 409. Weiss, W. P., H. R. Conrad, and W. L. Shockey. 1986. Amino acid profiles of heat-damaged grasses. J. Dairy Sci. 69: 1824. Weiss, W. P., H. R. Conrad, and W. L. Shockey. 1986. Digestibility of . nitrogen in heat-damaged alfalfa. J. Dairy Sci. 69: 2658. 198 TABLE 1. Mean percentage protein in each of 6 solubility fractions by forage species and maturity level across forage preservation methods (11 =12).l Crude protein fraction Forage 1 2 3 4 5 6 % Yld % of CP Alfalfa Early 39.3 3.9 5.8 22.5 3.6 18.9 93.9 Mid 36.6 4.1 5.1 20.7 4.7 24.9 96.1 Late 30.8 4.8 4.6 23.6 5.2 29.0 98.1 P, early vs, mid2 .05 NS NS .01 .001 .0001 NS P, mid vs. late .0001 .05 NS .0001 .1 .0001 NS Bromegrass Early 25.6 6.3 3.9 23.9 6.4 28.3 94.4 Mid 21.3 4.0 3.7 22.6 8.3 38.0 97.8 Late 17.6 3.1 4.2 21.9 8.5 40.5 - 96.0 P, early vs. mid .0001 .0001 NS .1 .0001 .0001 .01 P, mid vs. late .0001 .01 .01 NS NS .01 NS Canarygrass ' Early 28.6 5.3 3.4 23.8 7.0 25.4 93.5 Mid 27.3 5.2 4.0 20.2 7.4 26.0 91.4 Late 28.6 4.1 3.6 21.7 7.0 27.7 91.2 P, early vs. mid NS NS .05 .01 NS NS NS P, mid vs. late NS .1 NS NS NS .05 NS Corn Early 25.4 2.2 17.7 20.2 10.2 21.2 96.9 Late 24.7 2.6 21.5 21.0 8.4 19.5 97.8 P, early vs. late NS NS .001 NS NS .05 NS 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble 1n reduced med1a(b1carbonate), 3) soluble in 80% acetone, 4) soluble in pH-7 detergent at 95°C, 5) soluble 1n sulfuric ac1d and detergent at 95°C, and 6) insoluble. 2Significance of contrast, NS indicates P > .1. 199 TABLE 2. Mean protein percentage in each of 6 solubility fractions by forage species and preservation method across maturities (n = 9). Crude protein fraction Forage 1 2 3 4 5 6 % Yld % of CP Alfalfa Freeze dried 19.8 7.6 3.4 28.4 4.0 31.9 95.1 Hay 24.2 6.1 2.8 24.7 5.5 . 30.8 94.1 Silage (30% DM) 49.2 2.2 7.6 18.5 3.8 16.7 96.8 Silage (38% DM) 49.1 1.2 6.8 17.3 4.8 17.7 98.1 p, dried vs. ensiledz .0001 .001 .0001 .0001 .05 .(XJOI .05 P, freeze dried vs. .01 .01 NS .0001 .0001 NS NS hay P, silage 1 vs. 2 NS .05 NS .05 _.01 NS NS Bromegrass Freeze dried 10.0 9.0 2.6 31.8 7.3 35.5 96.1 Hay 15.0 6.4 3.0 24.1 8.8 40.0 97.3 Silage (33% DM) 29.7 1.4 5.9 17.6 7.6 30.0 92.1 Silage (44% DM) 31.4 1.1 4.2 17.7 7.2 37.0 98.7 P, dried vs. ensiled .0001 .0001 .0001 .0001 .05 .0001 NS P, freeze dried vs. .0001 .0001 .l .0001 .001 .0001 NS hay P, silage 1 vs. 2 .01 NS .0001 NS NS .0001 .0001 Canarygrass , - Freeze dried 12.8 6.9 3.3 33.0 9.4 25.8 91.1 Hay 20.9 7.1 2.6 20.0 6.2 34.0 90.9 Silage (34% DM) 42.8 2.0 4.6 18.3 6.3 22.3 96.1 Silage (43% DM) 36.1 3.4 4.2 16.4 6.6 23.4 90.1 P, dried vs. ensiled .0001 .0001 .0001 .0001 .001 .0001 NS P. freeze dried vs. .01 NS .05 .0001 .0001 .0001 NS hay P, silage 1 vs. 2 .05 1 NS NS NS NS .05 Corn Freeze dried 11.8 5.0 18.1 22.4 11.8 24.1 93.3 Silage (31% DM) 32.5 1.2 ' 20.2 19.2 8.2 18.3 99.5 Silage (38% DM) 30.8 1.1 20.5 20.2 8.1 18.5 99.2 P. dried vs. ensiled .0001 .0001 .05 .001 .01 .0001 .01 P, silage 1 vs. 2 .05 NS NS NS NS NS NS 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble in reduced media (bicarbonate), 3) soluble in 80% acetone, 4) soluble in pH-7 detergent at 95°C, 5) soluble in sulfuric acid and detergent at 95°C, and 6) insoluble. 2Significance of contrast, NS indicates P > .1. 200 TABLE 3. Prediction of forage protein degradation at 2 h determined from multiple regression analysis of protein fractions.1 Fraction Estimate S.E. P < ' Model 1 (R2 = .88) Intercept . 58.6 7.4 .0001 1 43.5 8.2 .0001 2 28.2 22.7 NS 3 -82.4 10.6 .0001 4 -31.4 11.1 .005 5 -103.6 21.3 .0001 6 -525 8.6 .0001 Model 2 (R2 = .82) 1 105.0 3.1 .0001 2 117.8 24.1 .0001 3 -31.0 10.3 .01 4 29.8 9.7 .01 5 -23.4 22.9 NS 6 2.6 6.1 NS Model 3 (R2 = .76) 1 + 2 103.1 3.3 .0001 3 ,t 4 1.5 5.5 NS 5 + 6 12.7 4.1 .05 Model 4 (R2 = .78) 1+2 996 29 .mmr 4 23.7 7.1 .001 6 3.3 5.3 NS 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble 1n reduced 1n vrtro med1a (bicarbonate), 3) soluble in 80% acetone, 4) soluble 1n neutral detergent at 95 C, 5) soluble in acid detergent at 95°C, and 6) insoluble. 2Probabi1ity that estimate at 0, NS indicates P > .1. 201 TABLE 4. Prediction of forage protein degradation at 24 h determined from multiple regression analysis of protein fractions.1 Fraction Estimate (%) S.E. P < Model 1 (R2 = .81) Intercept 62.1 8.1 .0001 1 32.7 9.0 .0005 2 3.4 24.8 NS 3 -73.2 11.6 .0001 4 5.7 12.1 NS 5 -138.3 23.2 .0001 6 42.4 9.3 .0001 Model 2' (R2 = .72) 1 98.8 3.3 .0001 2 98.5 26.1 .0002 3 -18.8 11.1 .1 4 70.6 10.5 .0001 5 -53.4 24.8 .05 6 15.9 6.6 .05 Model 3 (R2 = .54) 1 + 2 96.4 3.9 .0001 3+4 237 66 .mms 5 + 6 25.5 5.0 .0001 Model 4 (R2 = .67) 1 + 2 94.2 3.1 .0001 4 60.2 7.5 .0001 6 11.4 5.6 .05 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble in reduced media (bicarbonate), 3) soluble in 80% acetone, 4) soluble in pH-7 detergent at 95°C, 5) soluble in sulfuric acid and detergent at 95°C, and 6) insoluble. 2Probability that estimate a: 0, NS indicates p > .1. 202 TABLE 5. Percentage decrease from the initial amount of crude protein in solubility fractions after 2 h degradation in vitro. Negative values indicate percentage increase in crude protein in that fraction. Me 1 2 3 4 5 6 Alfalfa, early, freeze dried -46.6 83.8 -345.5 6.3 32.6 39.1 Alfalfa, mid, hay -24.0 81.8 -239.0 1.6 11.7 26.8 Bromegrass, late, hay -58.9 72.7 -l83.1 -14.2 1.6 34.5 Canarygrass, mid, freeze dried -43.2 74.9 -269.7 .4 3.0 35.9 Alfalfa, late, silage (%DM) 3.2 -19.4 -7l.5 8.1 6.0 -.7 Bromegrass, early, silage 1.6 ~3.9 -134.7 -4.8 ~17.8 31.9 Corn, late, silage (%DM) 6.7 .3 -21.1 -4.7 3.2 12.3 Canarygrass, mid, silage (%DM) 4.2 -31.6 ~100.7 -7.1 -28.4 32.2 SEM, n = 2 5.1 8.3 25.0 4.6 7.4 6.9 [3’ dried vs. silagez .0001 .0001 .0001 NS .01 .05 P, within dried .l .01 .07 .01 .05 NS P, within silage NS NS .06 NS 'Ns NS 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble in reduced media (bicarbonate), 3) soluble in 80% acetone, 4) soluble in pH-7 detergent at 95°C, 5) soluble in sulfuric acid and detergent at 95°C, and 6) insoluble. 2Significance of contrast, NS indicates P > .1. 203 TABLE 6. Percentage decrease from the initial amount of crude protein in the respective solublllty fractrons over time across forages. Negative values indicate percentage increase in crude protein in that fraction.1 0 Hi“ :~- -.~ 2mm nan—a a; .'_o‘ - , Time 1 2 3 4 5 6 Dried forages 2 h -43.2 78.3 -259.3 -1.5 12.2 34.1 6 h _ -46.4 83.3 -469.8 -3.3 54.7 34.6 24 h ~38.5 80.9 -539.2 -.1 37.6 ' 36.1 SEM, n = 8 4.5 1.5 24.1 1.5 3.2 3.3 Ensiled forages I l 2 h 3.9 -13.7 -82.0 -21 ' -9.3 18.9 6 h 11.7 29.1 -155.4 -3.5 17.2 4.9 24 h 7.9 2.5 -230.8 -8.7 30.4 8.0 SEM, n = 8 1.4 5.3 12.8 3.3 5.3 4.0 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble in reduced in vitro media (bicarbonate), 3) soluble in 80% acetone, 4) soluble 1n neutral detergent at 95°C, 5) soluble in acid detergent at 95°C, and 6) insoluble. 204 TABLE 7. Estimation using the Lucas test of percentage protein degraded from fractions. Slope of the line times 100 equals percentage of fraction degraded. The negative intercept results from metabolic protein, contamination from other protein fractions, or inappropriateness of the model.1 Estimate g 2 4 5 6 2 h degradation Intercept, % -1.9"'* -2.8 -2.7 -4.2"' Slope x 100 102.3” 11.6 25.6 48.8” R2 .99 .17 .06 .77 6 h degradation Intercept -1.1** -3.4 -6.3 -4.2 Slope x 100 97.5“ 12.6 86.7" 42.8” R2 .99 .1 1 .26 .54 24 h degradation . Intercept -l.7** -5.7* -4.5 -9.1** Slope x 100 102.5M 24.3* 70.4" 70.3" R2 .96 .31 .47 .83 *Probability that estimate at 0, P < .05, or ** P < .01. 1Crude protein fractions were extracted sequentially and separated by centrifugation as follows: 1) soluble in trichloroacetic acid, 2) soluble in reduced media (bicarbonate), 3) soluble in 80% acetone, 4) soluble in pH-7 detergent at 95°C, 5) soluble in sulfuric acid and detergent at 95°C, and 6) insoluble. 205 In vitro media (bicarbonate buffer) soluble insoluble Add trichloroacetic acid soluble 80% Acetone insoluble E ii I' l W soluble insoluble \ > E . 3 V Detergent, pH 7.0 soluble insoluble > E . I Detergent, 1 N H2804 soluble insoluble > E . 5 Fraction Figure 1. Scheme for sequential extraction of forage crude protein. 206 Figure 2. Prediction of forage protein degradation at 24 h from solubility fractions of original material. A: model included 6 fractions and intercept R2 = .81. B: model included fractions 1 + 2, 4 and 6, R2 = .67. nsof Predicted. protein degradation at 24 h (%) Predicted protein degradation at 24 h (%) 207 A. Model 1 70- 60: 50- 40- 30- 20- 10' 0 10 20 30 40 50 60 70 80 Measured protein degradation at 24 h (%) O B. Model 4 70- 60" 50‘ 40' 30- 20- 10‘ O 10 20 30 40 50 60 70 80 Measured protein degradation at 24 h (%) O i} 208 Figure 3. Change in crude protein solubility fractions due to degradation in ruminal enzyme extract. A: mean of 4 dried forages. B: mean of 4 ensiled forages. -- . --- Soluble in trichloroacetic acid, —- °-—- soluble in reduced media, -—l-- soluble in acetone, -- I —-- soluble in pH 7 detergent, -- Am soluble in l N H2804 with detergent, -- Am insoluble CP. 209 A. Dried Forages .5 to nw 2 50. 4o 30. 10. 0 EU .58 cc .5 5:25 5 £32.. 830 Time (h) B. Ensiled Forages 15 IO 50. 4o 30. 0 10. 0 Co .33 .e .5 =88... ... £38.. ..ch Time (h) 210 Figure 4. Lucas test for uniformity of degradation in ruminal enzymes across forages. 0 Indicates ensiled forages, I hay. Slope of the line is true degradation at time indicated. Intercept is "metabolic" protein. A: Buffer soluble CP (fraction 2) degraded at 2 h, y = 1.02x - 1.9, R2 = .99. B: Neutral detergent soluble CP (fraction 4) degraded at 24 h, y = .24x - 5.7, R2 = .31. 211 72 1'0 Initial Fraction 2 CP (% of total) l6 l4 - — _ - _ . 0 mol 8 6 4 2 O 7.. 2425 cc see N :2“an Soc m0 m0 304 15 - — _ - _ _ 0 93:5 mo 6ch V £030me 89¢ EU MO mmOsH Initial Fraction 4 CP (% of total) 212 Figure 5. Lucas test for uniformity of degradation in ruminal enzymes across forages. 0 Indicates ensiled forages, I hay. Slope of the line is true degradation at 24 h. Intercept is "metabolic" protein. A: Acid detergent soluble CP (fraction 5), y = .70x - 4.5, R2 = .47. B: Insoluble CP (fraction 6), y = .70x - 9.1, R2 = .83. 213 _1(). '2‘ o .2 8A 5-i I Era ge we 0- ° ° €361 ‘4-4 0 a ‘5 I I I I r 1 3 0 3 6 9 12 15 18 Initial Fraction 5 CP (% of total) B 20- \O t: 15- .2 8A 10" Era gzé 5‘ i3 0- 05° “5" -5" a '10 I I I I I r I I .3 0510152025 303540 Initial Fraction 6 CP (% of total) 214 CHAPTER 8: EPILOGUE The theory that high-producing ruminants respond to balancing rations for ruminally undegraded and degraded forms of protein is well established. However, the practice of balancing such rations is in its infantile stage. Previous research described in Chapter 1 has established that the amount of protein optimally degraded in the rumen varies depending production and nutritional factors. The amount of protein that is degraded in the rumen depends on animal and microbial factors, and on the protein source. The current methods to balance rations for degraded and undegraded protein do not begin to address this level of complexity. None-the-less, they do offer an opportunity to formulate and evaluate rations that generally meet the needs of ruminants. Research is needed to improve the estimation of protein requirements of cattle in order to reduce the amount of protein fed, and reduce costs, animal health and environmental concerns associated with feeding excess protein. Much of this needed research depends on better methods of chemical analysis of feeds. There is tremendous variation in protein degradation of forages (Chapters 1 and 6), including variation within a feed type, due to chemical or structural associations, supply of microbial growth or inhibition factors, or effectors of protease enzymes. Chemical analyses need to be deve10ped that can measure these modifiers of protein degradation in feeds. All ultimate goal is to evaluate a forage routinely for an individual producer, and determine the likely protein degradation when fed to a specific animal on a specific ration. Considering this goal, it is much more advantageous to determine likely rates of degradation of protein fractions than directly determining the extent. These rates can be 215 used to calculate the ruminal availability of protein by including the rates of passage into the model. Future work will require estimation of passage rates of fractions, and factors that affect them. Reduction of the problem of measuring ruminal. protein degradation to the two sub-problems, rate of degradation and rate of passage, allows for potentially greater understanding than could be obtained by trial and error evaluation of an infinite number of rations. Each of these sub-problems can be divided further into additional studies. Factors that affect protein degradation include the susceptibility of the feed to degradation, enzyme profiles in the rumen, and enzyme activity levels. The susceptibility of the feed to degradation was the part of the puzzle that was addressed directly in this dissertation. The hypothesis was that certain individual protein fractions separated based on solubility would degrade completely at uniform rates. If these fractions were uniform across forages, values for rates of degradation of the fractions could be compiled, and forages could be analyzed routinely for protein degradation pr0perties based on the composition of fractions. Further research could address factors that affect rates of degradation of fractions, and the susceptibility of fractions to be affected. The research presented in this dissertation identifies several additional problems that must be solved to analyze forages as described. The methods used to fractionate proteins require further deve10pment. The methods used to solubilize CP in buffers described in the Appendix and in Chapter 2 appeared to solubilize more true protein than the methods used in Chapter 7. This comparison is made across studies with different forages, but the differences might warrant further investigation. Differences in the methods used included use of fresh, undried material that was macerated in a tissue homogenizer for the first studies vs. use of freeze-dried material that was ground in a Wiley mill. In the studies on solubility described in the Appendix, even hay samples 216 which were ground in the Wiley mill were also macerated in the tissue homogenizer to be consistent with treatment of fresh forages. Another difference is that simpler buffers were used in earlier studies, but bicarbonate-based in vitro media was used with later studies. This difference was introduced so that solubility could be determined after in vitro degradation in the same buffer as used for degradation. A preliminary study did compare solubilization of alfalfa hay in phosphate buffer to in vitro media and showed no differences. Solubility of CP in buffers and TCA was also shown to increase with longer duration in the solvent, especially at higher incubation temperature (Appendix). Serine-protease inhibitors failed to prevent this solubilization in TCA, but other protease inhibitors may have been more successful and may be needed for these analyses. The means to preserve fresh forage samples before chemical analysis requires further attention. It was shown to be important not to allow samples to be frozen and thawed before analysis, and freeze drying was preferred (Chapter 2). However, further work will require comparison of fresh material to freeze dried samples to evaluate whether even this treatment is adequate. Addition of chemical reducing agents was needed to prevent precipitation of protein in buffers (Chapter 2). Methods to preserve samples on the farm for laboratory analysis will need to be developed. The laboratory fermentation method to measure protein degradation was labor intensive and had high variation within treatments. Some improvements in this method were made, and should be further validated. The importance of the method relates to validation of other procedures with a more physiological approach. The in vitro fermentation offers advantages over in situ fermentation in that total microbial growth can be evaluated, and end-products can be measured. The degradation of protein using crude extract from rumen contents offered several interesting observations. It appeared that proteolytic activity was resistant to self—degradation, and appeared to be auto- 217 activated. None-the—less, protein degradation rate was slow and not uniform within feeds. Rate of buffer solubilization was greater than TCA solubilization when degraded with enzyme, suggesting that proteins become filterable before complete degradation. Protein degradation of alfalfa hay was not dependent only on proteolytic activity, but responded to addition of cellulase. These observations suggest that part of the reason for the slow degradation of protein with enzyme extract compared to degradation by fermentation may have been due to absence of non-proteolytic enzymes that are required for degradation. If certain proteins depend on fiber digestion for their own degradation, then ruminal availability of these must take into account rate of degradation of fiber. This is another nutrient that degrades at differing rates across forages and is difficult to predict. These observations are important to consider when evaluating the degradation of protein fractions. While soluble protein was immediately degraded with enzyme extract, other fractions were degraded to differing extents within early points of the incubation, but degradation of some of these fractions slowed thereafter. The slowed degradation rate must have been due to lack of uniformity of the fractions with respect to degradation by rumen extract. Would these fractions have been uniform with respect to degradation by ruminal fermentation? The broader degradation by fermentation may have made parts of proteins available for degradation and prevented the slowing in degradation rate. In any case, the fractions selected were not uniform for degradation with enzyme, and so none of the fractions appeared to isolate only proteins that were dependent on non-proteolytic enzymes for their digestion. The problem of a lack of uniformity of protein fractions isolated can also be explained by the possibility that proteins cannot be separated by solubility into uniform fractions. It is possible that even a purified protein may degrade more slowly over time when in the presence of excess enzyme. The proteolytic enzymes cleave proteins 218 between certain residues as they become available. Might the number of available sites decline over time, and in fact might certain resistant peptides accumulate after degradation? If this is the case, isolating uniform fractions by solubility is not possible. The extent that this hypothesis is true if at all requires further investigation because it may be a severe limitation to development of solubility methods. The problem of predicting protein degradation of forages routinely in the laboratory is far from settled. This dissertation describes research that is more basic and systematic than many previous efforts that merely correlated simple laboratory methods with in situ data. Due to this approach, some gains have been made into understanding how proteins are degraded. Further efforts will have to address some of the new questions. What means of macerating and solubilizing proteins from forages are appropriate to isolate uniform fractions? How should forages be preserved before analysis? What are the structural limitations from non-protein forage components to degradation, and how are these measured? What is the nature of proteolytic enzymes in the rumen? Do individual proteins degrade completely at the same rate? The methods developed in this dissertation with further refinement might be used to answer some of these questions. 219 APPENDIX : PRELIMWARY EXPERIMENTS AND OBSERVATIONS Observations on Protein Solubility Measurement of Forage CP Solubility. An experiment using bovine albumen and bromegrass hay demonstrated that TCA was inefficient at precipitation of CP when samples were filtered through Whatman #54 filter paper instead of precipitated by centrifugation. The TCA soluble CP was higher for both samples, but more so for albumen. While the TCA addition to albumen solution turned the solution from clear to white, this protein was entirely filterable. Centrifugation speed (20,000 g vs. 1500 g) was compared for TCA precipitation of CP from bromegrass hay and albumen, with no difference between the two. Greater than 96% of the albumen was precipitated at either speed. There was no difference between centrifugation for 15 min or 30 min on CP precipitation of these same samples, and there was no improvement in precipitation by addition of deoxycholate (2 mM final concentration) to the solution. Lowry's method to determine CP in orchard grass was also attempted, but soluble CP was 45% higher when measured this way compared to Kjeldahl. Precipitation of protein and measurement of precipitate by Lowry's method resulted in 67% as much protein predicted compared to Kjeldahl. Interferences with Lowry's method appear to be abundant in forages. Bromegrass hay was solubilized in water (60 ml/ g DM) by adding water to sample at 25 °C and processing immediately. Samples (3 replicates of each) were strained through cheese cloth or filtered through Whatman #54 filter paper. The filtrate was sampled for N and the remaining sample was centrifuged at 1500 g for 30 min. ;— h _._.___ 220 True protein was precipitated with TCA (final conc. 6%, va01) and samples were centrifuged. Results are shown in Appendix Table 1. Filtering removed protein from solution compared to straining through cheese cloth, and centrifuging after filtering removed additional CP from the water. There was no effect of filtering before centrifuging, compared to straining before centrifuging. However, TCA precipitated less CP for filtered samples. Filtering was a slow process, and it was speculated that filtering may have provided time for plant proteolytic enzymes to degrade some water soluble protein to become TCA soluble. A subsequent experiment demonstrated an increase in TCA soluble CP in bromegrass by keeping it for 4 h in water at 25 °C vs. 15 min, but these samples were not different in buffer soluble CP. APPENDD( TABLE 1. Effect of straining through cheese cloth, filtering through Whatman # 54, or centrifuging (1500 g) on water soluble CP, or trichloroacetic acid (T CA) soluble CP in orchard grass hay. Treatment Soluble CP (% of DM) Filtered 5.7b Strained ~ 6.7a Filtered, centrifuged 4_9c Strained, centrifuged 4.5C Filtered, TCA precipitated 3.6d Strained, TCA precipitated 3.06 Standard Error of Mean (n = 3) .14 adeeMeans with different superscripts are significantly different, P < .05. Effect of pH and Eh on CP Solubility. An experiment was to evaluate effect of pH on buffer- and TCA-soluble protein of orchard grass hay. Soluble— and TCA—soluble crude protein in orchard grass bay (.5 g DM) were determined by storage with in vitro 221 media (30 ml) for 4 h at 25 °C under C02 pressure. The in vitro media was prepared with or without cysteine as a reducing agent (.6 m g/ml) and at 3 pH levels: 3, 7 and 10. Buffer solubility was determined by measurement of N in the supernatant after centrifuging (1500 g, 15 min) with subtraction of N from cysteine. In retrospect, sulfide would have been a better reducing agent. Solubility in TCA was determined by adding 10 m1 of 24% TCA (wt/vol) to sample that had been in 30 m1 of media for 4 h, and measuring N in supernatant after centrifuging. The pH and Eh of each tube were measured after centrifugation. Buffer pH and Eh (-340 to -250 mV) were maintained after soaking the sample. Buffer solubility increased at higher pH in unreduced samples, and solubility in both buffer and TCA seemed to increase with reduction, but since cysteine was used as reducing agent, this result may have occurred because of a mis—estimate of cysteine. The addition of TCA lowered pH to about 1.0 regardless of initial pH, but after addition of TCA, Eh ranged from +290 mV for unreduced samples and +180 mV for reduced samples. Buffering from media appeared to have no effect on TCA solubility. After addition of acid and efflux of C02, volume of media decreased, thus final concentration of TCA differed from expected by adding the sum of the parts. Higher buffer pH resulted in greater solubilization of true protein from orchard grass hay. Effect of Temperature on Proteolytic Activity of Alfalfa, Effect of soaking mid- maturity freeze-dried alfalfa (CP = 20.5%) in reduced in vitro media on TCA soluble CP was determined. Alfalfa samples (8 mg N) were soaked in media (30 ml) at 4°C or 38 °C in the absence or in presence of serine-protease inhibitor (3 mM phenylmethylsulfonyl fluoride, PMSF). These sammes were incubated for 3 h, but additional samples were incubated without protease inhibitor at 4 °C for 0 or 16 h. The results are shown in Appendix Table 2. There was no effect of adding PMSF to samples incubated at either 4 °C or 38 °C. However, samples stored at 4 °C 222 had lower TCA soluble CP than samples stored at 38 °C, and this value was not different from samples that were not stored in buffer (Treatment 5). Storage at 4 °C resulted in a slight increase in. TCA soluble CP at 16 h compared to 0 h. These results demonstrate that alfalfa samples may be soaked for up to 3 h at 4 °C without affecting the CP solubilized in TCA. However, soaking at 38 °C resulted in solubilization of CP in TCA which could be associated with plant proteases. However, these potential proteases were not inhibited by PMSF. APPENDD( TABLE 2. Effect of soaking one tenth bloom freeze-dried alfalfa on protein solubilized in TCA. Treatment Time Temperature Inhibitor % TCA sol. 1 3 h 4 °C PMSF 24,20 2 3 h 4 °C None 23_5C 3 3 h 38 °C PMSF 28.9a 4 3 h 38 °C None 30.021 5 0 h 4 °C None 24.2c 6 16 h 4 °c None 26.7b Means of each treatment are shown. Standard error of mean (n = 3) was .57. abCMeans with different superscripts differ using student T test (P < .05). Observations on In Vitro Fermentation Titration of in vitro media. Media (25 ml) used for in vitro rumen batch culture was titrated .to determine the amount of HCl required to lower pH to 4.0 in order to st0p activity after digestion and to remove H28 gas from the media. Addition of 3 m1 of 1 N HCl lowered pH to 2.5 with the pH dropping rapidly after reaching 6.0. However, over 223 time the pH increased again and 1 ml 6 N HCl is recommended to lower pH if this step is necessary. Variation in pH and reducing potential for in vitro batch culture. This experiment determined the stability in pH and reducing potential of the in vitro environment. Eight treatments (2 reps each) compared pH and Eh over time for different energy and protein contents. The treatments (expressed as g/tube) were low energy (.25 g cellulose, .25 g starch) and high energy (.5 g starch, .5 g cellulose) arranged factorially with no protein (.0156 g cysteine as only N), low protein (.0156 g cysteine + .0312 g casein) and high protein (.156 g cysteine + .0625 g casein). An additional treatment included 1 g of each starch and cellulose and .0625 g protein. APPENDIX TABLE 3. Effect of starch and cellulose level (E) and casein level (P) on pH and Eh of in vitro batch culture. Imment 0.11 2h 21h 48h 12h l.QO_h pH Eh Eh pH Eh pH Eh pH Eh pH Eh low E no P ~289 6.8 ~318 6.7 ~281 6.3 ~210 6.1 ~141 ~279 6.7 ~329 6.7 ~288 6.3 ~l99 6.2 ~178 lowElowP 6.9 ~278 ~248 6.6 ~382 6.2 283 6.1 ~353 5.9 ~188 6.9 ~286 ~290 6.2 ~363 6.1 ~242 5.9 ~237 5.9 ~175 lowEhiP 6.8 -240 ~l90 6.4 ~312 6.0 ~l94 5.7 ~102 ~~ ~- 6.8 ~272 ~280 6.4 ~341 6.1 ~24] 5.8 ~167 5.9 ~127 hiEnoP ~280 6.6 ~325 6.5 ~275 6.1 ~l91 5.8 ~l67 ~285 6.5 ~341 6.5 ~324 6.0 ~224 5.8 ~174 hiElowP ~280 6.3 343 5.7 197 5.5 147 5.5 ~135 289 6.3 ~311 5.8 ~182 5.5 ~139 5.4 ~l30 hiEhiP ~265 6.3 -326 5.7 ~183 5.4 ~102 5.4 ~77 ~25] 6.3 ~33? 5.8 ~l93 5.5 ~135 5.4 ~85 veryhiEhiP ~250 5.6 ~433 5.3 ~307 5.0 ~l81 5.0 ~119 ~290 5.9 ~333 5.5 ~247 5.3 ~157 5.2 ~129 224 Both energy and protein level affected pH and Eh despite the fact that energy was supplied in great excess of that of protein (Appendix Table 3). This suggests that adding a lot of extra energy will change the fermentation. Simply oversupplying energy is not appropriate. Instead, an amount of starch or fiber should be added so that tubes testing both forages and concentrates have the same amount of each energy source. The amount of cysteine added may need to be reduced and the total amount of protein and energy should be at least at the low values shown here. Effect of Reducing Agents on Microbial Fermentation, An experiment was conducted to determine the effect of ascorbate on microbial growth in vitro. Media (30 ml) containing soybean meal (100 mg), starch (200 mg) and cellulose (160 mg) was incubated for 3, 8 or 24 h with rumen inocula (8 ml). Media (30 ml) was previously reduced with sodium salt of ascorbate (0.4 g/L) and cysteine (0.04 g/L), and was labeled at reduction with 353-cysteine (5.0 x 105 dpm). APPENDIX TABLE 4. Effect of sodium ascorbate addition on precipitation of 358 (dpm) with soybean meal fermented with rumen microorganisms. Ascorbate Time Rep 1 Rep} Rep 3 No 3 h 31,447 32,596 23,617 8 h 41,404 49,872 50,426 24 h 49,872 42,511 35,021 Yes 3 h 14,766 16,468 17,617 8 h 30,851 31,149 32,553 24 h 34,638 36,638 41,191 Precipitable 35S (dpm) is given in Appendix Table 4. The level of N in the pellet did not change consistently due to the treatment but was greater for 8 and 24 h than for 3 h (data not shown). Protein degradation was not calculated because 35S:N 225 ratios of microbial protein could not be determined. The TCA precipitated 358 increased over time and was reduced by addition of sodium ascorbate. These results imply that addition of ascorbate inhibited microbial growth and protein degradation (since precipitated N was constant). Effect of carbohydrate source on microbial growth and 3SS:N ratio. Several carbohydrates were compared for their ability to support microbial growth on trypticase (36 mg). Treatments were corn starch (200 mg), corn starch (400 mg), gelatinized corn starch (200 mg DM), cellulose (160 mg) wheat straw neutral detergent fiber (W SNDF) (200 mg), starch and cellulose, and no carbohydrate. Gelatinized starch was prepared by heating 10% starch solution for 50 min until it became viscous. Samples were incubated for 22 h in media containing rumen inocula, 35S-cysteine (8.9 x 105 dpm) and cysteine (lg/1). Samples were rinsed three times in solution containing 1% HBr and 1 g/L of each cysteine.HCI, Na2804 and NaSO3. Radioactivity in the pellet was determined after digestion in H2804 with H202. The pH was reduced from 6.85 to 6.65 for gelatinized starch compared to non-gelatinized at 4 h of incubation. It was reduced from 6.54 to 6.40 at 22 h of digestion for gelatinized and not gelatinized starch receptively. The results in Appendix Table 5 indicate that the presence of starch in the in vitro media is important to support microbial growth. The various starch treatments were not different from one another in precipitable N. There was no difference in microbial N precipitation between cellulose, no energy and wheat straw NDF. The dpm:N ratio was much higher for WS NDF suggesting that the precipitation of non- microbial 358 would be a bigger problem with some feeds than others. The level of unlabeled sulfur would have to be decreased to allow prediction of precipitable microbial N from precipitable S. 226 APPENDD( TABLE 5. Precipitation of 35S after incubation of 358-cysteine with rumen microorganisms, trypticase and the carbohydrate sources listed as treatments. Treatment dpm N (mg) dpm:N No addition 22,3974,b 1.5a 15,904a 200 mg starch 32,7801),C 3.1b 10,629a 400 mg starch 40,3120 35*) . 11,750a Gelatinized starch 31,901b.c 3.4b 9,404a Cellulose 15,7 30a 1.2a 13,047a Starch + Cellulose 39,248c 3.1b 10,868a ws NDF 55,078d 1.8a 30,860b SEM (n = 3) 3284 0.2 2,089 abcMeans with different superscripts differ (P<.05). Observations on Crude Rumen Fluid Extract Reducing agents on protein degradation. Sodium sulphide (1.2 g/l), cysteine.HCl (1.2 g/l), a combination thereof (.6 gll each) and no reduction were compared for soybean meal degradation at 16 h. All samples were incubated with perfusion of C02 over samples in the presence of crude extract from rumen fluid. Samples were accidentally incubated at 30 °C instead of 38 °C. Appendix Table 6 shows that, though reducing agents are necessary, either sulfide, cysteine.HCl or both were equally adequate. These results contrast those from azocascin degradation, which showed cysteine to result in higher proteolytic activity than sulfide, because the method used for feed degradation may not allow sulfide to escape as easily due to the larger volume of media. However, if C02 is diffused over the samples for the entire time it might result in a loss of reducing potential from evolution of H28. 227 Alfalfa hay was degraded using several methods of reducing the sample and using other variations. In vitro media (10 ml) was reduced with .6 gll of each NaS and cysteine.HCl except where indicated otherwise. Feeds were incubated with 0, 5 or 10 ml of enzyme for 16 h and compared to 0 or 5 ml of enzyme that was immediately precipitated with feed. One treatment differed in that it was presoaked with reduced media for 2 h before addition of enzyme and 16 h incubation. Variation in reduction of media included no reduction, saturation with C02 by dispersion of gas in media for 30 min after reduction (allows removal of some H28 gas), use of 25 ml media instead of 10 ml, and use of 3 gll of each reducing agent instead of .6 gll. Each treatment was replicated twice. ' APPENDIX TABLE 6. Effect of reductants on soybean meal (SBM) degradation with enzyme extracted from the rumen. Reduction SBM degradation (%) No reduction 6.4b Na2S (1.2 gll) 20.8a Cysteine.HCl (1.2 g/l) 23.1a Na28 + Cysteine.HCl (.6 gll each) 243a aDifferent superscripts indicate differences among means, 11 = 3, P < .05). Results of the experiment on reduction and other treatments of luceme are difficult to interpret owing to the small sample size. These results are shown in Appendix Table 7. Since there were only two replicates per treatment raw data are shown, but specific contrasts were made with certain combined treatments. Comparison of treatments 1~2 with 4-5 shows that alfalfa does in fact appear to degrade over time by the plant's own proteases (P > .0001). One may interpret this increase in 228 APPENDIX TABLE 7. Effect of the presence of enzyme and reducing method on protein degradation of alfalfa. % Alfalfa degraded Treat— Enzyme Time (h) Media Presoak Reduce Rep 1 Rep 2 ment (ml) (ml) 1 0 16 10 no yes 51.1 45.1 2 5 16 10 no yes 44 44.3 3 10 16 10 no yes 44.3 44.2 4 0 0 10 no yes 30.7 32.3 5 5 0 10 no yes 30.6 30 6 5 16 10 yes yes 48 46.7 7 0 16 10 no no 45.4 41.7 8 5 16 10 no no 42.4 42.6 9 0 16 10 no before 41.0 41.7 C02 10 5 16 10 no before 42.8 45.8 C02 1 1 0 16 25 no yes 47.0 46.8 12 5 16 25 no yes 44.5 46.5 13 0 16 10 no 3 gll 40.2 42.2 14 5 16 10 no 3 gll 46.5 46.2 229 degradation as the leaching of NPN out of the luceme forage. However, in another experiment where alfalfa was soaked in the refrigerator, the increase in TCA soluble protein was not observed. Contrast of treatments 1, 4, 7, 9, 11 and 13 with treatments 2, 5, 8, 10, 12 and 14 shows the effect of adding the ruminal enzyme, and it was not significant P > .1). This degradation in the absence of enzymes was not found for SBM. It leaves questions about the value of extracting ruminal enzyme for forage degradation. Soybean meal protein did not degrade to any appreciable extent without addition of enzyme, indicating that the enzyme does in fact degrade feed proteins. Contrast of slightly reduced treatments, 7-10 with highly reduced treatments 11-14 was significant (P < .02) indicating that careful reduction can affect luceme protein degradation. Centrifugation of rumen extract before use. Effect of centrifuging enzyme extract after thawing and before incubation with enzyme on CP degradation of soybean meal and alfalfa hay. Thawed enzyme extract (5 ml) was centrifuged at 26,000 g before adding to soybean meal samples, or it was not centrifuged. The experiment was repeated twice with enzyme extract from different days with two replicates per day. The first time, 2.65 mg N precipitated after 16 h in samples incubated without feeds (blanks) for uncentrifuged enzyme extract compared to non-detectable levels for centrifuged enzyme. Soybean meal was 45.6% TCA soluble after 16 h degradation for samples incubated with uncentrifuged enzyme compared to 54.3 % for centrifuged enzyme. The second time, blank N was .75 mg for uncentrifuged vs. non-detectable levels for centrifuged extract, and soybean meal was 54.2 vs. 58.8 for uncentrifuged and centrifuged enzyme extract respectively. A different study measured the proteolytic activity on azocascin on both centrifuged and uncentrifuged samples of enzyme extract. There was no loss of proteolytic activity from centrifugation an additional time, but the change in the amount of blank N indicates that the interference is greatly reduced. 230 Higher predictions of protein degradation are observed, especially when blank N is high, for centrifuged enzyme. These observations led to centrifuging enzyme before use for all experiments. I Presoaking of Alfalfa. An experiment with two replicates of each treatment per run for two runs, showed that alfalfa hay CP was more soluble in TCA for samples that were soaked in cold buffer for 2 h before addition of enzyme extract for 2 h than for samples that were added directly to enzyme extract. In this experiment, there was no change in TCA soluble CP from 0 to 2 h for the samples that were not presoaked, but 11.5 % of initially insoluble CP was solubilized for presoaked samples. There was no significant effect of the same treatments on soybean meal. Since there was a reduced lag when samples were presoaked in cold media before addition of enzyme extract, and since there was negligible degradation of CP by plant enzymes at 4 °C, presoaking at 4 °C was adapted at a treatment for all feeds subjected to degradation with crude rumen extract. Tungstic acid vs. TCA precipitation. An experiment was conducted to compare tungstic acid vs. TCA precipitation of corn and soybean meal protein and enzyme extract. Treatments included soybean meal or corn grain (8 mg N) or no added feed incubated for 0 h or 16 h with 10 ml ruminal extract prepared as described in Chapter 5, except it was not centrifuged before use. Each feed was also precipitated immediately without addition of enzyme, and each treatment was replicated twice. Tungstic acid precipitated 70% of the enzyme preparation before incubation compared to 42% for TCA (Appendix Table 8). Corn protein appeared poorly degraded when TCA was used, and actually appeared to increase when tungstic acid precipitation was used. Feeds precipitated without extract and without incubation resulted in greater precipitation of feed than that calculated from subtracting blanks 231 from feeds precipitated with enzyme (Appendix Table 9). Enzyme extract or feed CP appeared to precipitate to a lessor extent when together than when separate. 232 APPENDIX TABLE 8. Precipitation of CP before and after 16 h incubation with crude rumen extract using trichloroacetic acid (TCA) or tungstic acid (W). 0 h Feed TCA w TCA w mg N Enzyme only 5.6b 9.5b 2.0c 5.4b SBM w/ enzyme 11.19 15.08 6.0b 10.1a Com w/ enzyme 11831 14.2a 7.7a 11.5a SEM (n = 3) .35 .39 .12 .40 % of Total N Precipitated Enzyme only 41.83 70.1%! 14.79 39.851 SBM w/ enzyme 51.7a 69.93 28.0b . 47.0a Com w/ enzyme 54.7a 658%! 35.7a 53.3a SEM (n = 3) 4.2 4.1 1.1 4.4 % of 0 h Precipitated Enzyme only 35.2b 56.8c SBM w/ enzyme --- --- 54.2a 67.3b Com w/ enzyme 65.33 81.03 SEM (n = 2) -- 3.2 2.1 ab‘3Means with different superscripts differ P < .05. 233 APPENDD( TABLE 9. Interference of crude rumen extract with measurement of precrpltable protein (mg N of total 8 mg) in corn or soybean meal (SBM) before degradation. Feed TCA Tungstic acid Corn with rumen extract 6.1 4.7 Corn alone 6.8 6.3 SBM with rumen extract 5.5 5.6 SBM alone 6.6 6.3 SEM (n = 2) .35 .39 1Feed protein was precipitated from buffer that either contained rumen enzyme extract or did not contain extract. Feed protein precipitated was calculated as total CP precipitated minus precipitated CP in media with only extract for samples combined with extract. Main effect of presence of extract was significant P < .05. II HUI/I21! LIBRQRIES W) I ”thrust. ». . «in HICHIGRN STATE UNIV . . Iii? stun...“ .42.... . . .2 Kill... . art. . a}. rMHFEt ... ...“... I» ... 4.111.? ...i. 1531 I 4.34..