@article{heugten_martinez_mccomb_koutsos_2022, title={Improvements in Performance of Nursery Pigs Provided with Supplemental Oil Derived from Black Soldier Fly (Hermetia illucens) Larvae}, volume={12}, ISSN={["2076-2615"]}, DOI={10.3390/ani12233251}, abstractNote={The current study evaluated the impact of increasing levels of supplemental black soldier fly larvae (BSFL) oil, a novel and sustainable feed ingredient, on growth performance and blood chemistry indices in nursery pigs. Pigs were weaned at 21 days of age (n = 192; body weight = 6.9 ± 0.15 kg) and randomly allotted within sex and body weight to 1 of 4 dietary treatments, using 48 pens (4 pigs/pen). Treatments contained 0, 2, 4, or 6% supplemental BSFL oil, replacing equivalent quantities of corn oil. A 3-phase feeding program was used with 14, 11, and 15 days for phase 1 to 3, respectively. Supplementation of BSFL oil linearly (p ≤ 0.052) improved pig body weight and daily gain throughout the study and overall. No differences were observed in feed intake; however, feed efficiency was linearly improved with BSFL oil supplementation for phase 1 and 2 (p < 0.05). Serum cholesterol increased linearly (p = 0.002) and platelet count tended to increase linearly (p = 0.082) with increasing BSFL oil. No other differences were detected in blood chemistry and all results were within normal ranges. In conclusion, BSFL oil is a valuable, energy-dense feed ingredient that can promote growth performance of nursery pigs.}, number={23}, journal={ANIMALS}, author={Heugten, Eric and Martinez, Gabriela and McComb, Alejandra and Koutsos, Liz}, year={2022}, month={Dec} } @article{anderson_holt_boyd_heugten_2021, title={Impact of Replacing Soybean Meal with Corn DDGS and Crystalline Amino Acids on Performance and Carcass Characteristics of Growing Pigs}, volume={99}, ISSN={["1525-3163"]}, DOI={10.1093/jas/skab054.119}, abstractNote={Abstract}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Anderson, Brooke E. and Holt, Jon P. and Boyd, R. D. and Heugten, Eric}, year={2021}, month={May}, pages={72–72} } @article{padilla_jha_fellner_van heugten_2021, title={PSIII-12 In vitro Evaluation of Purified Fiber Sources for Production of Short-chain Fatty Acids Using Pig Cecal Content as an Inoculum}, volume={99}, ISSN={0021-8812 1525-3163}, url={http://dx.doi.org/10.1093/jas/skab054.299}, DOI={10.1093/jas/skab054.299}, abstractNote={Abstract}, number={Supplement_1}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Padilla, Gabriela E Martinez and Jha, Rajesh and Fellner, Vivek and van Heugten, Eric}, year={2021}, month={May}, pages={177–177} } @article{silva-guillen_arellano_martinez_heugten_2020, title={Growth performance, oxidative stress, and antioxidant capacity of newly weaned piglets fed dietary peroxidized lipids with vitamin E or phytogenic compounds in drinking water}, volume={36}, ISSN={["2590-2865"]}, url={http://dx.doi.org/10.15232/aas.2019-01976}, DOI={10.15232/aas.2019-01976}, abstractNote={ABSTRACT Objective This study evaluated the use of vitamin E and phytogenic compounds in drinking water on growth performance, oxidative stress, and immune status of piglets fed peroxidized lipids. Materials and Methods In a 35-d study, 21-d-old weaned piglets (n = 96; 6.10 ± 0.64 kg of BW) were assigned within sex and BW blocks to 1 of 4 treatments, using 24 pens (4 pigs per pen; 6 replications per treatment). Diets contained either 6% soybean oil or 6% peroxidized soybean oil. Pigs fed peroxidized soybean oil received drinking water without (control) or with supplemental vitamin E (100 IU/L of RRR-α-tocopherol) or phytogenic compounds (60 μL/L for wk 1 and 30 μL/L for wk 2 to 5). Results and Discussion Peroxidized soybean oil decreased (P Implications and Applications Peroxidized soybean oil reduced growth performance of weaned nursery pigs, which did not appear to be related to oxidative stress or immune status. The negative effects of peroxidized soybean oil on animal performance could not be improved by supplementation of vitamin E or phytogenic compounds in the drinking water.}, number={3}, journal={APPLIED ANIMAL SCIENCE}, publisher={American Registry of Professional Animal Scientists}, author={Silva-Guillen, Ysenia and Arellano, Consuelo and Martinez, Gabriela and Heugten, Eric}, year={2020}, month={Jun}, pages={341–351} } @article{mendoza_boyd_zier-rush_ferket_haydon_heugten_2017, title={Effect of natural betaine and ractopamine HCl on whole-body and carcass growth in pigs housed under high ambient temperatures}, volume={95}, DOI={10.2527/jas2017.1622}, number={7}, journal={Journal of Animal Science}, author={Mendoza, S. M. and Boyd, R. D. and Zier-Rush, C. E. and Ferket, P. R. and Haydon, K. D. and Heugten, E.}, year={2017}, pages={3047–3056} } @article{shah_lentz_van heugten_currin_singletary_2017, title={Tempering ventilation air in a swine finishing barn with a low-cost earth-to-water heat exchanger}, volume={9}, ISSN={1941-7012}, url={http://dx.doi.org/10.1063/1.4979359}, DOI={10.1063/1.4979359}, abstractNote={An earth-to-water heat exchanger (EWHE) can reduce livestock heat stress and also save electricity and water. A 4-kW EWHE system comprising 154 m of a polyvinyl chloride (PVC) pipe (35 mm ID) buried in 3.2 m of soil was evaluated for its ability to provide cooling to 60 pigs in a finishing barn in Raleigh, NC. A low-cost tube-and-fin heat exchanger was used to exchange energy between water (38 l/min) and air (0.58 to 1.22 m3/s). After 8 h of cooling, at 1.22 m3/s, the temperature change (ΔT), energy produced (qh), and coefficient of performance (COP) were as high as 3 °C, 4.3 kW, and 8.2, respectively. After 12 h of continuous operation for air tempering during winter, |ΔT|, |qh|, and COP were 2.2 °C, 3.4 kW, and 6.7, respectively. While the EWHE pens were slightly warmer than the Control pens cooled with stir fans and sprinklers on very hot days, pig performance in the EWHE pens was unaffected. The EWHE reduced the electricity use by >50% and eliminated the sprinkling water use. Burying plastic pipes in slinky coils instead of using double pass rigid PVC pipes could improve system performance as would wetting the soil around the pipes. In addition to being sustainable, the EWHE could be cost-effective for zone-cooling of high-value pigs as well as greenhouse cooling in many parts of the world.}, number={2}, journal={Journal of Renewable and Sustainable Energy}, publisher={AIP Publishing}, author={Shah, Sanjay B. and Lentz, Zachary A. and van Heugten, Eric and Currin, Richard D., Jr. and Singletary, Isaac}, year={2017}, month={Mar}, pages={023901} } @article{rosero_boyd_mcculley_odle_heugten_2016, title={Essential fatty acid supplementation during lactation is required to maximize the subsequent reproductive performance of the modern sow}, volume={168}, ISSN={["1873-2232"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84979483427&partnerID=MN8TOARS}, DOI={10.1016/j.anireprosci.2016.03.010}, abstractNote={This study was conducted to investigate the effects of supplemental essential fatty acids (EFA) on sow reproductive efficiency and to estimate the concentrations of EFA required by the lactating sow for maximum subsequent reproduction. Data were collected on 480 sows (PIC Camborough) balanced by parity, with 241 and 239 sows representing Parity 1, and 3-5 (P3+), respectively. Sows were assigned randomly, within parity, to a 3 × 3 factorial arrangement plus a control diet without added lipids. Factors included linoleic (2.1%, 2.7%, and 3.3%) and α-linolenic acid (0.15%, 0.30%, and 0.45%), obtained by adding 4% of different mixtures of canola, corn and flaxseed oils to diets. Diets were corn-soybean meal based with 12% wheat middlings. The benefits of supplemental EFA were more evident for the subsequent reproduction of mature P3+ sows. For these sows, supplemental α-linolenic acid improved the proportion of sows that farrowed relative to sows weaned (linear P=0.080; 82.8, 80.5, and 92.8% for sows fed 0.15%, 0.30%, and 0.45% α-linolenic acid, respectively). In addition, supplemental linoleic acid, fed to Parity 1 and P3+ sows, tended to increase subsequent litter size (linear P=0.074; 13.2, 13.8 and 14.0 total pigs born for 2.1%, 2.7% and 3.3% linoleic acid, respectively). These results demonstrate that a minimum dietary intake of both α-linolenic and linoleic acid is required for the modern lactating sow to achieve a maximum reproductive outcome through multiple mechanisms that include rapid return to estrus, increased maintenance of pregnancy and improved subsequent litter size.}, journal={ANIMAL REPRODUCTION SCIENCE}, publisher={Elsevier BV}, author={Rosero, David S. and Boyd, R. Dean and McCulley, Mark and Odle, Jack and Heugten, Eric}, year={2016}, month={May}, pages={151–163} } @misc{rosero_boyd_odle_heugten_2016, title={Optimizing dietary lipid use to improve essential fatty acid status and reproductive performance of the modern lactating sow: a review}, volume={7}, ISSN={["2049-1891"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-85008336841&partnerID=MN8TOARS}, DOI={10.1186/s40104-016-0092-x}, abstractNote={Dietary lipid supplementation benefits the prolific and high-producing modern lactating sow. A comprehensive review of recent studies showed that lipid supplementation increases average daily energy intake, which is partitioned for lactation as indicated by greater milk fat output and improved litter growth rate. Recent compelling findings showed that addition of particular lipids during lactation improved the subsequent reproductive outcome of sows. Such benefits were related to the level of dietary essential fatty acids (EFA, linoleic acid, C18:2n-6; and α-linolenic acid, C18:3n-3) during lactation. Lactation diets without supplemental EFA resulted in a pronounced negative balance (intake minus milk output) of linoleic (−25.49 g/d) and α-linolenic acid (−2.75 g/d); which compromised sow fertility (farrowing rate < 75 % and culling rates > 25 % of weaned sows). This phenomenon seems to be increasingly important with advancing sow age because of a progressive reduction of body EFA pool over successive lactations. The net effect of supplemental EFA during lactation was to create a positive EFA balance, which improved the subsequent reproduction of sows. Adequate linoleic acid intake improved the proportion of sows that farrowed in the subsequent cycle (Farrowing rate (%) = [(−1.5 × 10−3 × linoleic acid intake (g/d)2) + (0.53 × linoleic acid intake (g/d)) + (45.2)]; quadratic P = 0.002, R2 = 0.997, RMSE = 0.031). In addition, increasing linoleic acid intake increased the number of pigs born in the subsequent cycle (total pigs born (n) = [(9.4 × 10−5 × linoleic acid intake (g/d)2) + (0.04 × linoleic acid intake (g/d)) + (10.94)]; quadratic P = 0.002, R2 = 0.997, RMSE = 0.031). Supplemental α-linolenic acid resulted in a rapid return to estrus (sows bred: sows weaned = 94.2 %; wean-to-estrus interval = 4.0 d) and achieved a high retention of pregnancy (sows pregnant: sows bred = 98 %). Collectively, we conclude that a minimum dietary intake of 10 g/d of α-linolenic acid, simultaneous with a minimum of 125 g/d of linoleic acid should be provided to ≥ 95 % of the sows; thereby, achieving a maximum sow reproductive efficiency through multiple mechanisms that include rapid return to estrus, high maintenance of pregnancy and large subsequent litter size in mature sows, that appear to be susceptible to EFA deficiency.}, number={1}, journal={JOURNAL OF ANIMAL SCIENCE AND BIOTECHNOLOGY}, author={Rosero, David S. and Boyd, R. Dean and Odle, Jack and Heugten, Eric}, year={2016}, month={Jun} } @article{adeola_azain_carter_crenshaw_estienne_kerr_lindemann_maxwell_miller_shannon_et al._2015, title={A cooperative study on the standardized total-tract digestible phosphorus requirement of twenty-kilogram pigs}, volume={93}, ISSN={["1525-3163"]}, DOI={10.2527/jas.2015-9509}, abstractNote={A cooperative study comprising growth performance, bone mineralization, and nutrient balance experiments was conducted at 11 stations to determine the standardized total-tract digestible (STTD) P requirement of 20-kg pigs using broken-line regression analysis. Monocalcium phosphate and limestone were added to a corn-soybean meal-based diet at the expense of cornstarch to establish 6 concentrations of STTD P from 1.54 to 5.15 g/kg in increments of 0.62 g/kg at a constant Ca:total P of 1.52:1.0. Diets were fed to 936 pigs (average initial BW of 19 kg) in 240 pens for 20 replicate pens of barrows and 20 replicate pens of gilts per diet. As STTD P increased from 1.54 to 5.15 g/kg of the diet for d 0 to 14, 14 to 28, and 0 to 28, the ADG, ADFI, and G:F increased ( < 0.01). Barrows gained and ate more ( < 0.05) than gilts during d 14 to 28 and 0 to 28. There was no interaction between sex and STTD P concentration for any of the growth performance response criteria. There were both linear and quadratic increases ( < 0.05) in mineral density and content of ash, Ca, and P in the femur expressed as a percentage of dry, fat-free metacarpal as dietary STTD P increased. Furthermore, the maximum load of the femur and mineral density and content and maximum load as well as the Ca and P expressed as a percentage of metacarpal ash linearly increased ( < 0.01) with increasing dietary concentrations of STTD P. There were both linear and quadratic increases ( < 0.01) in apparent digestibility and retention of P with increasing concentrations of STTD P in the diets. Digestibility and retention of Ca linearly ( < 0.01) increased with increasing dietary concentrations of STTD P. Breakpoints determined from nonlinear broken-line regression analyses revealed estimates of 4.20 ± 0.102, 3.20 ± 0.036, or 3.87 ± 0.090 g/kg for ADG during d 0 to 14, 14 to 28, or 0 to 28, respectively. Corresponding estimates using G:F as the response criterion were 4.34 ± 0.146, 3.38 ± 0.139, or 4.08 ± 0.195 g/kg. When mineralization of the femur was used as criteria of response, estimates of STTD P requirement were 4.28, 4.28, or 4.34, g/kg for mineral density, mineral content, or maximum load, respectively. Using mineralization of the metacarpal as criteria of response, estimates of STTD P requirement ranged from 3.5 to 5.0 g/kg depending on the metacarpal response criteria. The study provided empirical estimates of STTD P requirements of 20- to 40-kg pigs.}, number={12}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Adeola, O. and Azain, M. J. and Carter, S. D. and Crenshaw, T. D. and Estienne, M. J. and Kerr, B. J. and Lindemann, M. D. and Maxwell, C. V. and Miller, P. S. and Shannon, M. C. and et al.}, year={2015}, month={Dec}, pages={5743–5753} } @article{rosero_odle_arellano_boyd_heugten_2015, title={Development of prediction equations to estimate the apparent digestible energy content of lipids when fed to lactating sows}, volume={93}, ISSN={["1525-3163"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84965190994&partnerID=MN8TOARS}, DOI={10.2527/jas.2014-8402}, abstractNote={Two studies were conducted 1) to determine the effects of free fatty acid (FFA) concentrations and the degree of saturation of lipids (unsaturated to saturated fatty acids ratio [U:S]) on apparent total tract digestibility (ATTD) and DE content of lipids and 2) to derive prediction equations to estimate the DE content of lipids when added to lactating sow diets. In Exp. 1, 85 lactating sows were assigned randomly to a 4 × 5 factorial arrangement of treatments plus a control diet with no added lipid. Factors included 1) FFA concentrations of 0, 18, 36, and 54% and 2) U:S of 2.0, 2.8, 3.5, 4.2, and 4.9. Diets were corn-soybean meal based and lipid was supplemented at 6%. Concentrations of FFA and U:S were obtained by blending 4 lipid sources: choice white grease (CWG; FFA = 0.3% and U:S = 2.0), soybean oil (FFA = 0.1% and U:S = 5.5), CWG acid oil (FFA = 57.8% and U:S = 2.1), and soybean-cottonseed acid oil (FFA = 67.5% and U:S = 3.8). Titanium dioxide was added to diets (0.5%) as a digestibility marker. Treatments started on d 4 of lactation and fecal samples were collected after 6 d of adaptation to diets on a daily basis from d 10 to 13. The ATTD of added lipid and DE content of lipids were negatively affected (linear, < 0.001) with increasing FFA concentrations, but negative effects were less pronounced with increasing U:S (interaction, < 0.05). Coefficients of ATTD for the added lipid and DE content of lipids increased with increasing U:S (quadratic, = 0.001), but these improvements were less pronounced when the FFA concentration was less than 36%. Digestible energy content of added lipid was described by DE (kcal/kg) = [8,381 - (80.6 × FFA) + (0.4 × FFA) + (248.8 × U:S) - (28.1 × U:S) + (12.8 × FFA × U:S)] ( = 0.74). This prediction equation was validated in Exp. 2, in which 24 lactating sows were fed diets supplemented with 6% of either an animal-vegetable blend (A-V; FFA = 14.5% and U:S = 2.3) or CWG (FFA = 3.7% and U:S = 1.5) plus a control diet with no added lipids. Digestible energy content of A-V (8,317 and 8,127 kcal/kg for measured and predicted values, respectively) and CWG (8,452 and 8,468 kcal/kg for measured and predicted values, respectively) were accurately estimated using the proposed equation. The proposed equation involving FFA concentration and U:S resulted in highly accurate estimations of DE content (relative error, +0.2 to -2.3%) of commercial sources of lipids for lactating sows.}, number={3}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Rosero, D. S. and Odle, J. and Arellano, C. and Boyd, R. D. and Heugten, E.}, year={2015}, month={Mar}, pages={1165–1176} } @article{rosero_odle_mendoza_boyd_fellner_heugten_2015, title={Impact of dietary lipids on sow milk composition and balance of essential fatty acids during lactation in prolific sows}, volume={93}, ISSN={["1525-3163"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84965094122&partnerID=MN8TOARS}, DOI={10.2527/jas.2014-8529}, abstractNote={Two studies were designed to determine the effects of supplementing diets with lipid sources of EFA (linoleic and α-linolenic acid) on sow milk composition to estimate the balance of EFA for sows nursing large litters. In Exp. 1, 30 sows, equally balanced by parity (1 and 3 to 5) and nursing 12 pigs, were fed diets supplemented with 6% animal-vegetable blend (A-V), 6% choice white grease (CWG), or a control diet without added lipid. Diets were corn-soybean meal based with 8% corn distiller dried grains with solubles and 6% wheat middlings and contained 3.25 g standardized ileal digestible Lys/Mcal ME. Sows fed lipid-supplemented diets secreted greater amounts of fat (P = 0.082; 499 and 559 g/d for control and lipid-added diets, respectively) than sows fed the control diet. The balance of EFA was computed as apparent ileal digestible intake of EFA minus the outflow of EFA in milk. For sows fed the control diet, the amount of linoleic acid secreted in milk was greater than the amount consumed, throughout lactation. This resulted in a pronounced negative balance of linoleic acid (-22.4, -38.0, and -14.1 g/d for d 3, 10, and 17 of lactation, respectively). In Exp. 2, 50 sows, equally balanced by parity and nursing 12 pigs, were randomly assigned to a 2 × 2 factorial arrangement of diets plus a control diet without added lipids. Factors included linoleic acid (2.1% and 3.3%) and α-linolenic acid (0.15% and 0.45%). The different concentrations of EFA were obtained by adding 4% of different mixtures of canola, corn, and flaxseed oils to diets. The n-6 to n-3 fatty acid ratios in the diets ranged from 5 to 22. Increasing supplemental EFA increased (P < 0.001) milk concentrations of linoleic (16.7% and 20.8%, for 2.1% and 3.3% linoleic acid, respectively) and α-linolenic acid (P < 0.001; 1.1 and 1.9% for 0.15 and 0.45% α-linolenic acid, respectively). Increasing supplemental EFA increased the estimated balance of α-linolenic acid (P < 0.001; -0.2 and 5.3 g/d for 0.15% and 0.45% α-linolenic acid, respectively), but not linoleic acid (P = 0.14; -3.4 and 10.0 g/d for 2.1% and 3.3% linoleic acid, respectively). In conclusion, lipid supplementation to sow lactation diets improved milk fat secretion. The fatty acid composition of milk fat reflected the dietary supplementation of EFA. The net effect of supplemental EFA was to create a positive balance during lactation, which may prove to be beneficial for the development of nursing piglets and the subsequent reproduction of sows.}, number={6}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Rosero, D. S. and Odle, J. and Mendoza, S. M. and Boyd, R. D. and Fellner, V. and Heugten, E.}, year={2015}, month={Jun}, pages={2935–2947} } @article{rosero_odle_moeser_boyd_heugten_2015, title={Peroxidised dietary lipids impair intestinal function and morphology of the small intestine villi of nursery pigs in a dose-dependent manner}, volume={114}, ISSN={["1475-2662"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84949192029&partnerID=MN8TOARS}, DOI={10.1017/s000711451500392x}, abstractNote={Abstract}, number={12}, journal={BRITISH JOURNAL OF NUTRITION}, author={Rosero, David S. and Odle, Jack and Moeser, Adam J. and Boyd, R. Dean and Heugten, Eric}, year={2015}, month={Dec}, pages={1985–1992} } @article{mendoza_heugten_2014, title={Effects of dietary lipid sources on performance and apparent total tract digestibility of lipids and energy when fed to nursery pigs}, volume={92}, ISSN={["1525-3163"]}, DOI={10.2527/jas.2013-6488}, abstractNote={Acidulated fats and oils are by-products of the fat-refining industry. They contain high levels of FFA and are 10% to 20% less expensive than refined fats and oils. Two studies were designed to measure the effects of dietary lipid sources low or high in FFA on growth performance and apparent total tract digestibility (ATTD) of lipids and GE in nursery pigs. In Exp. 1, 189 pigs at 14 d postweaning (BW of 9.32 ± 0.11 kg) were used for 21 d with 9 replicate pens per treatment and 3 pigs per pen. Dietary treatments consisted of a control diet without added lipids and 6 diets with 6% inclusion of lipids. Four lipid sources were combined to create the dietary treatments with 2 levels of FFA (0.40% or 54.0%) and 3 degrees of fat saturation (iodine value [IV] = 77, 100, or 123) in a 2 × 3 factorial arrangement. Lipid sources were soybean oil (0.3% FFA and IV = 129.4), soybean-cottonseed acid oil blend (70.5% FFA and IV = 112.9), choice white grease (0.6% FFA and IV = 74.8), and choice white acid grease (56.0% FFA and IV = 79.0). Addition of lipid sources decreased ADFI (810 vs. 872 g/d; P = 0.018) and improved G:F (716 vs. 646 g/kg; P < 0.001). Diets high in FFA tended (P = 0.08) to improve final BW (21.35 vs. 21.01 kg) and ADG (576 vs. 560 g/d). Lipid-supplemented diets had greater ATTD of lipids than control diets (67.4% vs. 29.7%; P < 0.001). Apparent total tract digestibility of lipids was greater in diets with low FFA (69.9% vs. 64.9%; P < 0.001) and decreased linearly with increasing IV (73.2%, 69.1%, and 67.2%). For GE, ATTD was greater in diets with low FFA (83.1% vs. 80.9%; P = 0.001). In Exp. 2, 252 pigs at 7 d postweaning (BW of 7.0 ± 0.2 kg) were used for 28 d with 9 replicate pens per treatment and 4 pigs per pen. Diets included a control diet without added lipids and 6 treatments with 2.5%, 5.0%, or 7.5% of lipids from either poultry fat (1.9% FFA) or acidulated poultry fat (37.8% FFA) in a 2 × 3 factorial arrangement. Addition of lipids increased (P < 0.001) final BW (19.9 vs. 18.4 kg) and ADG (460 vs. 405 g/d) regardless of source. Fat increased (P < 0.001) ADFI when added at 2.5% and then decreased ADFI with each further increment (663, 740, 681, and 653 g for 0%, 2.5%, 5.0%, and 7.5% fat, respectively). Inclusion of lipids linearly (P < 0.001) improved G:F (615, 615, 688, and 692 g/kg for 0%, 2.5%, 5.0%, and 7.5% fat, respectively) and ATTD of lipids (17.8%, 50.2%, 71.0%, and 77.3% for 0, 2.5, 5.0, and 7.5% fat, respectively) and GE (76.1%, 76.4%, 83.3%, and 84.4% for 0%, 2.5%, 5.0%, and 7.5% fat, respectively). Acidulated lipids resulted in similar performance compared with refined lipids and could be economical alternatives to more expensive lipid sources.}, number={2}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Mendoza, S. M. and Heugten, E.}, year={2014}, month={Feb}, pages={627–636} } @article{dal jang_lindemann_heugten_jones_kim_maxwell_radcliffe_2014, title={Effects of phytase supplementation on reproductive performance, apparent total tract digestibility of Ca and P and bone characteristics in gestating and lactating sows}, volume={27}, number={3}, journal={Revista Colombiana de Ciencias Pecuarias}, author={Dal Jang, Y. and Lindemann, M. D. and Heugten, E. and Jones, R. D. and Kim, B. G. and Maxwell, C. V. and Radcliffe, J. S.}, year={2014}, pages={178–193} } @article{koger_o'brien_burnette_kai_kempen_heugten_kempen_2014, title={Manure belts for harvesting urine and feces separately and improving air quality in swine facilities}, volume={162}, ISSN={["1878-0490"]}, DOI={10.1016/j.livsci.2014.01.013}, abstractNote={Modern swine facilities have not been designed to maximize manure value nor to minimize NH3 emission. These benefits can possibly be achieved by harvesting urine and feces separately using a conveyor belt placed at a 4° angle beneath the slats. Urine drains from this belt into a gutter leading to a closed storage vessel while feces remain on the belt for up to 24 h. Such a belt was evaluated in a partially slatted swine facility housing 80–100 grower pigs in five separate experiments. Fecal DM was determined as a function of both belt residence time and collection time-of-day. The driest feces were obtained with daily collections at 0600 h. Collections at this time of day resulted in a 9.8±5.0% increase in DM over collection at 1500 h (P=0.07). Under steady state conditions, feces were collected at 49±5% DM and output was 0.26±0.05 kg DM pig−1 d−1 suggesting an apparent feed DM digestibility of 82.8±2.1%. Urine collected was 1.3±0.2 L pig−1 d−1, equivalent to 33±6% of the water intake. Emissions from this facility were for ammonia 1.03±0.20 kg pig−1 yr−1 or 5.9±1.0% of the intake N and for methane 1.05±0.29 kg pig−1 yr−1 or 0.64±0.18% of the feed energy. Odor emission at the ventilation fan was 1.9 OU animal−1 s−1. All three emission parameters were substantially less than literature values for conventional houses. In conclusion, the belt system was easy to operate and allowed for the separate collection of urine and feces resulting in reduced odor nuisance. Technically, it resulted in feces that could be harvested at 49% DM, and emissions of only 1 kg NH3 and CH4 pig−1 yr−1.}, journal={LIVESTOCK SCIENCE}, author={Koger, J. B. and O'Brien, B. K. and Burnette, R. P. and Kai, R. and Kempen, M. H. J. G. and Heugten, E. and Kempen, T. A. T. G.}, year={2014}, month={Apr}, pages={214–222} } @article{price_lin_van heugten_odle_willis_odle_2013, title={Diet physical form, fatty acid chain length, and emulsification alter fat utilization and growth of newly weaned pigs}, volume={91}, ISSN={0021-8812 1525-3163}, url={http://dx.doi.org/10.2527/jas.2012-5307}, DOI={10.2527/jas.2012-5307}, abstractNote={An experiment was conducted to examine the interplay of diet physical form (liquid vs. dry), fatty acid chain length [medium- (MCT) vs. long-chain triglyceride (LCT)], and emulsification as determinants of fat utilization and growth of newly weaned pigs. Ninety-six pigs were weaned at 20.0 ± 0.3 d of age (6.80 ± 0.04 kg) and fed ad libitum 1 of 8 diets for 14 d according to a 2(3) factorial arrangement of treatments with 6 pens per diet and 2 pigs per pen. The MCT contained primarily C8:0 and C10:0 fatty acids, whereas the LCT mainly contained C16:0, C18:0, C18:1, and C18:2. Diet physical form greatly impacted piglet growth (P < 0.001), with liquid-fed pigs (486 g/d) growing faster than dry-fed pigs (332 g/d) by 46%. Pigs fed LCT grew 22% faster (P = 0.01) than MCT-fed pigs; however, effects of emulsifier were not detected (P > 0.1). Furthermore, feed intake and G:F were 15% and 29% greater for liquid-fed pigs, and intake also was 21% greater for pigs fed LCT (P = 0.01). Diet physical form had no effect on apparent ileal fatty acid digestibility, but as expected, digestibility was greater (P < 0.001) for the MCT than the LCT diet (98.5% vs. 93.4%). Emulsification improved digestibility of most fatty acids in pigs fed LCT but not MCT (interaction, P < 0.01). Both jejunal and ileal villi height increased from 7 to 14 d postweaning (P < 0.01). Liquid-fed pigs had greater jejunal crypt depth (P < 0.05) compared with pigs fed the dry diet; however, ileal morphology was not affected by diet physical form, fat chain length, or emulsification. Plasma ketone body concentrations were 6-fold greater in pigs fed MCT than LCT, and the difference was greater in pigs fed dry diets (interaction, P = 0.01). The bile salt concentration in jejunal digesta was 2.2-fold greater in pigs fed LCT than in pigs fed MCT (P < 0.001). Collectively, we conclude that feeding liquid diets containing emulsified LCT can improve fat utilization and markedly accentuate feed intake, growth, and G:F of weanling pigs.}, number={2}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Price, K. L. and Lin, X. and van Heugten, E. and Odle, R. and Willis, G. and Odle, J.}, year={2013}, month={Feb}, pages={783–792} } @article{herfel_jacobi_lin_van heugten_fellner_odle_2013, title={Stabilized rice bran improves weaning pig performance via a prebiotic mechanism}, volume={91}, ISSN={0021-8812 1525-3163}, url={http://dx.doi.org/10.2527/jas.2012-5287}, DOI={10.2527/jas.2012-5287}, abstractNote={Stabilized rice bran (SRB) is classified as a "functional food" because of its prebiotic characteristics. With increasing grain prices and the pressure to remove antibiotics from swine diets because of concern over antibiotic resistance, SRB was investigated as a nursery diet ingredient with and without the addition of antibiotics (ANT). Two hundred pigs were weaned at 21 d of age, blocked by BW, and allotted to diets containing 0 or 10% SRB ± ANT according to a 2 × 2 factorial arrangement of treatments. Five animals were housed per pen throughout a 28-d growth period. At the end of the trial, 1 pig from each pen was euthanized for measurement of intestinal morphology. Antibiotic supplementation improved ADG by 6.4% during Phase 2 (d 14 to 28; P = 0.02), but other production variables were unaffected by ANT. During Phase 2 and cumulatively (d 0 to 28), the supplementation of SRB improved G:F by 10% in ANT-free pigs but not in pigs fed ANT (ANT × SRB, P < 0.03). Ileal histology revealed an increase in crypt depth of pigs fed the diet containing ANT plus SRB and corresponding decreases in villi:crypt associated with both ANT and SRB supplementation (P < 0.05). Intraepithelial lymphocytes were increased by 15% in pigs fed SRB without ANT, but were unaffected by SRB in pigs fed ANT (ANT x SRB, P = 0.003). Colonic bifidobacteria tended to increase with SRB supplementation (P < 0.10). Differences in ileal and cecal digesta short-chain fatty acid concentrations were not detected. In summary, SRB improved the efficiency of nutrient utilization in nursery diets lacking antibiotics and tended to increase intestinal bifidobacteria concentrations, indicating that SRB may exert beneficial prebiotic effects in weanling pigs.}, number={2}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Herfel, T. and Jacobi, S. and Lin, X. and Van Heugten, E. and Fellner, V. and Odle, J.}, year={2013}, month={Feb}, pages={907–913} } @article{shields_heugten_odle_stark_2012, title={Impact of crude glycerol on feed milling characteristics of swine diets}, volume={175}, ISSN={["0377-8401"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84863434334&partnerID=MN8TOARS}, DOI={10.1016/j.anifeedsci.2012.05.008}, abstractNote={Abstract This study evaluated the effect of crude glycerol on feed processing of nursery and finisher diets for pigs. In Exp. 1, complex nursery pig diets were mixed in a double ribbon mixer with glycerol levels of 0, 25, and 50 g/kg replacing maize on a weight basis and pelleted (74 °C conditioning temperature (CT) and 4.4 mm × 29 mm die size). In Exp. 2, finisher diets (maize–soybean meal based with 30 g/kg added fat) were mixed in a double ribbon mixer with glycerol included at 0, 25, and 50 g/kg replacing maize on a weight basis and pelleted (74 °C CT; 4.4 mm × 29 mm die size). A second control feed was pelleted at a standard CT of 85 °C. There were three replications per treatment for Exp. 1 and 2. In Exp. 1, flowability of mash nursery diets improved linearly (P}, number={3-4}, journal={ANIMAL FEED SCIENCE AND TECHNOLOGY}, author={Shields, M. C. and Heugten, E. and Odle, J. and Stark, C. S.}, year={2012}, month={Aug}, pages={193–197} } @article{sullivan_freeman_van heugten_ange-van heugten_wolfe_poore_2012, title={Impact of two types of complete pelleted, wild ungulate feeds and two pelleted feed to hay ratios on the development of urolithogenic compounds in meat goats as a model for giraffes}, volume={97}, ISSN={0931-2439}, url={http://dx.doi.org/10.1111/j.1439-0396.2012.01297.x}, DOI={10.1111/j.1439-0396.2012.01297.x}, abstractNote={Summary}, number={3}, journal={Journal of Animal Physiology and Animal Nutrition}, publisher={Wiley}, author={Sullivan, K. and Freeman, S. and van Heugten, E. and Ange-van Heugten, K. and Wolfe, B. and Poore, M. H.}, year={2012}, month={Apr}, pages={566–576} } @article{rosero_heugten_odle_arellano_boyd_2012, title={Response of the modern lactating sow and progeny to source and level of supplemental dietary fat during high ambient temperatures}, volume={90}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84882715021&partnerID=MN8TOARS}, DOI={10.2527/jas.2012-4242}, abstractNote={The objective of this study was to determine the response to increments of 2 sources of dietary fat on lactating sow and progeny performance during high ambient temperatures. Data were collected from 391 sows (PIC Camborough) from June to September in a 2,600-sow commercial unit in Oklahoma. Sows were randomly assigned to a 2 × 3 factorial arrangement of treatments and a control diet. Factors included 1) fat sources, animal-vegetable blend (A-V) and choice white grease (CWG), and 2) fat levels (2%, 4%, and 6%). The A-V blend contained 14.5% FFA with an iodine value of 89, peroxide value of 4.2 mEq/kg, and anisidine value of 23, whereas CWG contained 3.7% FFA with an iodine value of 62, peroxide value of 9.8 mEq/kg, and anisidine value of 5. Diets were corn-soybean meal based, with 8.0% distillers dried grains with solubles and 6.0% wheat middlings, and contained 3.56-g standardized ileal digestible Lys/Mcal ME. Sows were balanced by parity, with 192 and 199 sows representing parity 1 and parity 3 to 5, respectively. Feed refusal increased linearly (P < 0.001) with the addition of supplemental fat, but feed and energy intake increased linearly (P < 0.01) with increasing dietary fat. Sows fed CWG diets had reduced (linear, P < 0.05) BW loss during lactation. Litter growth rate was not affected by additional dietary fat. Addition of CWG to the diets improved G:F (sow and litter gain relative to feed intake) compared with the G:F of sows fed the control diet or the diets containing the A-V blend (0.50, 0.43, and 0.44, respectively; P < 0.05). Gain:ME (kg/Mcal ME) was greater (P < 0.05) for CWG (0.146) than A-V blend (0.129) but was not different from that of the control diet (0.131). Addition of A-V blend and CWG both improved (P < 0.05) conception and farrowing rates and subsequent litter size compared with the control diet. In conclusion, energy intake increased with the addition of fat. The A-V blend contained a greater amount of aldehydes (quantified by anisidine value) and was more susceptible to oxidation, resulting in reduced feed efficiency than CWG. Subsequent litter size and reproductive performance was improved by inclusion of both sources of fat in diets fed to lactating sows.}, number={8}, journal={Journal of Animal Science}, author={Rosero, D. S. and Heugten, E. and Odle, Jack and Arellano, C. and Boyd, R. D.}, year={2012}, pages={2609–2619} } @article{rosero_heugten_odle_cabrera_arellano_boyd_2012, title={Sow and litter response to supplemental dietary fat in lactation diets during high ambient temperatures}, volume={90}, ISSN={["0021-8812"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-84856294956&partnerID=MN8TOARS}, DOI={10.2527/jas.2011-4049}, abstractNote={The objective of this experiment was to determine the impact of supplemental dietary fat on total lactation energy intake and sow and litter performance during high ambient temperatures (27 ± 3°C). Data were collected from 337 mixed-parity sows from July to September in a 2,600-sow commercial unit in Oklahoma. Diets were corn-soybean meal-based with 7.5% corn distillers dried grains with solubles and 6.0% wheat middlings and contained 3.24 g of standardized ileal digestible Lys/Mcal of ME. Animal-vegetable fat blend (A-V) was supplemented at 0, 2, 4, or 6%. Sows were balanced by parity, with 113, 109, and 115 sows representing parity 1, 2, and 3 to 7 (P3+), respectively. Feed disappearance (subset of 190 sows; 4.08, 4.18, 4.44, and 4.34 kg/d, for 0, 2, 4, and 6%, respectively; P < 0.05) and apparent caloric intake (12.83, 13.54, 14.78, and 14.89 Mcal of ME/d, respectively; P < 0.001) increased linearly with increasing dietary fat. Gain:feed (sow and litter BW gain relative to feed intake) was not affected (P = 0.56), but gain:Mcal ME declined linearly with the addition of A-V (0.16, 0.15, 0.15, and 0.14 for 0, 2, 4, and 6%, respectively; P < 0.01). Parity 1 sows (3.95 kg/d) had less (P < 0.05) feed disappearance than P2 (4.48 kg/d) and P3+ (4.34 kg/d) sows. Body weight change in P1 sows was greater (P < 0.01) than either P2 or P3+ sows (-0.32 vs. -0.07 and 0.12 kg/d), whereas backfat loss was less (P < 0.05) and loin depth gain was greater (P < 0.05) in P3+ sows compared with P1 and P2 sows. Dietary A-V improved litter ADG (P < 0.05; 1.95, 2.13, 2.07, and 2.31 kg/d for 0, 2, 4, and 6% fat, respectively) only in P3+ sows. Sows bred within 8 d after weaning (58.3, 72.0, 70.2, and 74.7% for 0, 2, 4, and 6%, respectively); conception rate (78.5, 89.5, 89.2, and 85.7%) and farrowing rate (71.4, 81.4, 85.5, and 78.6%) were improved (P < 0.01) by additional A-V, but weaning-to-breeding interval was not affected. Rectal and skin temperature and respiration rate of sows were greater (P < 0.002) when measured at wk 3 compared with wk 1 of lactation, but were not affected by A-V addition. Parity 3+ sows had lower (P < 0.05) rectal temperature than P1 and P2 sows, and respiration rate was reduced (P < 0.001) in P1 sows compared with P2 and P3+ sows. In conclusion, A-V improved feed disappearance and caloric intake, resulting in improved litter weight gain and subsequent reproductive performance of sows; however, feed and caloric efficiency were negatively affected.}, number={2}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Rosero, D. S. and Heugten, E. and Odle, J. and Cabrera, R. and Arellano, C. and Boyd, R. D.}, year={2012}, month={Feb}, pages={550–559} } @article{graves_liwimbi_israel_heugten_robinson_cahoon_lubbers_2011, title={Distribution of ten antibiotic resistance genes in E. coli isolates from swine manure, lagoon effluent and soil collected from a lagoon waste application field}, volume={56}, ISSN={["0015-5632"]}, DOI={10.1007/s12223-011-0019-z}, abstractNote={The prevalence of ten antibiotic resistance genes (ARGs) was evaluated in a total of 616 Escherichia coli isolates from swine manure, swine lagoon effluent, and from soils that received lagoon effluent on a commercial swine farm site in Sampson County, North Carolina (USA). Isolates with ARGs coding for streptomycin/spectinomycin (aadA/strA and strB), tetracycline (tetA and tetB), and sulfonamide (sul1) occurred most frequently (60.6-91.3%). The occurrence of E. coli isolates that carried aadA, tetA, tetB, and tetC genes was significantly more frequent in soil samples (34.0-97.2%) than in isolates from lagoon samples (20.9-90.6%). Furthermore, the frequency of isolates that contain genes coding for aadA and tetB was significantly greater in soil samples (82.6-97.2%) when compared to swine manure (16.8-86.1%). Isolates from the lagoon that carried tetA, tetC, and sul3 genes were significantly more prevalent during spring (63.3-96.7%) than during winter (13.1-67.8%). The prevalence of isolates from the lagoon that possessed the strA, strB, and sul1 resistance genes was significantly more frequent during the summer (90.0-100%) than during spring (66.6-80.0%). The data suggest that conditions in the lagoon, soil, and manure may have an impact on the occurrence of E. coli isolates with specific ARGs. Seasonal variables seem to impact the recovery isolates with ARGs; however, ARG distribution may be associated with mobile genetic elements or a reflection of the initial numbers of resistant isolates shed by the animals.}, number={2}, journal={FOLIA MICROBIOLOGICA}, author={Graves, A. K. and Liwimbi, L. and Israel, D. W. and Heugten, E. and Robinson, B. and Cahoon, C. W. and Lubbers, J. F.}, year={2011}, month={Mar}, pages={131–137} } @article{shields_van heugten_lin_odle_stark_2011, title={Evaluation of the nutritional value of glycerol for nursery pigs}, volume={89}, ISSN={0021-8812 1525-3163}, url={http://dx.doi.org/10.2527/jas.2010-3558}, DOI={10.2527/jas.2010-3558}, abstractNote={In Exp. 1, a total of 144 pigs (BW, 6.68 ± 0.17 kg) were weaned at 21 d, blocked by BW, and allocated to 48 pens with 3 pigs per pen. Pens were randomly assigned to 1 of 6 dietary treatments (0, 2.5, 5, 7.5, and 10% glycerol supplemented to replace up to 10% lactose in a basal starter 1 diet containing 20% total lactose, which was fed for 2 wk), and a negative control diet with 10% lactose and 0% glycerol. A common starter diet was fed for the next 2 wk. In Exp. 2, a total of 126 pigs (BW, 6.91 ± 0.18 kg) were weaned at 21 d of age, blocked by BW, and allocated to 42 pens with 3 pigs per pen. Pigs were assigned to 1 of 6 treatments in a 2 × 3 factorial arrangement in a randomized complete block design with factors being 1) glycerol inclusion in replacement of lactose in starter 1 diets (0 or 5%) fed for 2 wk, and 2) glycerol inclusion in starter 2 diets (0, 5, or 10%) fed for 3 wk. In Exp. 1, glycerol supplementation at 10% improved (P=0.01) ADG (266 vs. 191 g/d) and G:F (871 vs. 679 g/kg) during the starter 1 period when compared with the negative control. Incremental amounts of glycerol linearly (P<0.05) increased ADG and ADFI, but did not affect G:F during starter 1. There was no effect of feeding glycerol during the starter 1 phase on subsequent performance during the starter 2 phase or overall. Serum glycerol concentrations increased linearly (P=0.003) with increasing dietary glycerol, and serum creatinine (P=0.004) and bilirubin (P=0.03) concentrations decreased with increasing glycerol. In Exp. 2, glycerol did not affect performance during starter 1, but it linearly increased (P≤0.01) ADG and ADFI during starter 2 (464, 509, and 542 and 726, 822, and 832 g/d, respectively) and overall (368, 396, and 411 and 546, 601, and 609 g/d, respectively). At the end of the study, pigs were 1.0 and 1.5 kg heavier when fed 5 and 10% glycerol, respectively (linear, P<0.01). Serum glycerol concentrations increased linearly during starter 2 (P<0.001), but were not affected during starter 1. Glycerol supplementation increased serum urea N quadratically (P<0.001) and decreased creatinine linearly (P<0.05) in the starter 2 phase. Overall, data indicate that glycerol can be added to nursery pig diets at 10%, while improving growth performance.}, number={7}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Shields, M. C. and van Heugten, E. and Lin, X. and Odle, J. and Stark, C. S.}, year={2011}, month={Jul}, pages={2145–2153} } @article{holt_heugten_graves_see_morrow_2011, title={Growth performance and antibiotic tolerance patterns of nursery and finishing pigs fed growth-promoting levels of antibiotics}, volume={136}, ISSN={["1878-0490"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-79951956142&partnerID=MN8TOARS}, DOI={10.1016/j.livsci.2010.09.007}, abstractNote={The objective of this study was to evaluate the impact of growth-promoting levels of antibiotics in diets for nursery and finishing pigs on growth performance and antibiotic tolerance patterns. Gilts (n = 200, initial body weight was 6.2 ± 0.003 kg), were allotted based on body weight to one of four treatments in a 2 × 2 factorial randomized complete block design. Nursery treatments consisted of feeding no antibiotics (CON) or an antibiotic diet (ANTI) containing chlortetracycline (CTC; 55 mg/kg). At the end of the nursery phase, one-half of the pigs receiving CON were switched to a diet containing antibiotic (virginiamycin; VIR, 11 mg/kg) and one-half of the pigs receiving ANTI were switched to CON for the remainder of the trial. This created four treatments for the finishing phase, consisting of: control in nursery and finishing (CC), antibiotic in nursery, control in finishing (AC), control in nursery, antibiotic in finishing (CA), or antibiotics throughout (AT). The pigs were weighed at the diet changes during the nursery (weeks 1, 3, and 5) and finishing (weeks 7, 9, 13, 17, and 20) phases. Fecal samples were collected at all diet changes for isolation of fecal coliforms and Enterococcus and subsequently tested for tolerance to CTC and VIR. After 1 week, CON pigs weighed less (7.09 vs. 7.28 kg) and had lower ADG (149 vs. 180 g/day) and ADFI (174 vs. 192 g/day) than ANTI pigs (P < 0.05). No performance differences were observed during the remainder of the study. At the initiation of the study (week 0), the ability of coliforms to grow in the presence of CTC and VIR, respectively, were 68 and 73% and increased to 90 and 96% at week 19 (time effect, P < 0.001). At week 17, tolerance of coliforms to CTC was greater for CA (98%) than AC (86%, time × treatment effect, P < 0.004). Enterococcus tolerance to CTC at week 7 was lower for CC (55%) compared to AT (76%), AC (74%) and CA (83%, time × treatment effect, P < 0.001). At week 9, Enterococcus tolerant to CTC and VIR, respectively, was lower for CC (15 and 18%) than AT (31 and 40%), AC (35 and 35%), and CA (44 and 43%, time × treatment effect, P < 0.001). Antibiotic growth promoters had little impact on growth performance in clean, isolated facilities with high labor inputs. The tolerance of bacteria to antibiotics fluctuated over time and persisted regardless of the use of antibiotic growth promoters.}, number={2-3}, journal={LIVESTOCK SCIENCE}, author={Holt, J. P. and Heugten, E. and Graves, A. K. and See, M. T. and Morrow, W. E. M.}, year={2011}, month={Apr}, pages={184–191} } @misc{chaytor_hansen_heugten_see_kim_2011, title={Occurrence and Decontamination of Mycotoxins in Swine Feed}, volume={24}, ISSN={["1976-5517"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-79960243133&partnerID=MN8TOARS}, DOI={10.5713/ajas.2011.10358}, abstractNote={Contamination of agricultural crops by mycotoxins results in significant economic losses for grain producers and, when consumed, it can cause reduced growth and health in a wide range of animal species. Hundreds of mycotoxin producing molds exist, however each has a different frequency and pattern of occurrence, as well as differences in the severity of the diseases (mycotoxicoses) they cause. Among the mycotoxins considered to be major contaminates are aflatoxin, deoxynivalenol, fumonisin, ochratoxin, and zearalenone. Although a multitude of species can be harmed by consumption of these mycotoxins, swine appear to be the most commonly affected commodity species. The swine industry can thus experience great losses due to the presence of mycotoxin contamination in feeds. Subsequently, recognition and prevention of mycotoxicoses is extremely important and dependent on adequate grain sampling and analysis methods pre-harvest, as well as effective strategies post-harvest to reduce consumption by animals. The aim of this review is to provide an overview of the major mycotoxin contaminants in grains, to describe methods of analysis and prevention to reduce mycotoxicoses in swine and other animals, and finally to discuss how mycotoxins directly affect swine production. (}, number={5}, journal={ASIAN-AUSTRALASIAN JOURNAL OF ANIMAL SCIENCES}, author={Chaytor, Alexandra C. and Hansen, Jeff A. and Heugten, Eric and See, M. Todd and Kim, Sung Woo}, year={2011}, month={May}, pages={723–738} } @article{bruininx_borne_heugten_milgen_verstegen_gerrits_2011, title={Oxidation of Dietary Stearic, Oleic, and Linoleic Acids in Growing Pigs Follows a Biphasic Pattern}, volume={141}, ISSN={["0022-3166"]}, DOI={10.3945/jn.111.142562}, abstractNote={We used the pig as a model to assess the effects of dietary fat content and composition on nutrient oxidation and energy partitioning in positive energy balance. Pigs weighing 25 kg were assigned to either: 1) a low fat-high starch diet, or 2) a high saturated-fat diet, or 3) a high unsaturated-fat diet. In the high-fat treatments, 20% starch was iso-energetically replaced by 10.8% lard or 10.2% soybean oil, respectively. For 7 d, pigs were fed twice daily at a rate of 1200 kJ digestible energy · kg(-0.75) · d(-1). Oral bolus doses of [U-(13)C] glucose, [U-(13)C] α-linoleate, [U-(13)C] stearate, and [U-(13)C] oleate were administered on d 1, 2, 4, and 6, respectively, and (13)CO(2) production was measured. Protein and fat deposition were measured for 7 d. Fractional oxidation of fatty acids from the low-fat diet was lower than from the high-fat diets. Within diets, the saturated [U-(13)C] stearate was oxidized less than the unsaturated [U-(13)C] oleate and [U-(13)C] linoleate. For the high unsaturated-fat diet, oxidation of [U-(13)C] oleate was higher than that of [U-(13)C] linoleate. In general, recovery of (13)CO(2) from labeled fatty acids rose within 2 h after ingestion but peaked around the next meal. This peak was induced by an increased energy expenditure that was likely related to increased eating activity. In conclusion, oxidation of dietary fatty acids in growing pigs depends on the inclusion level and composition of dietary fat. Moreover, our data suggest that the most recently ingested fatty acids are preferred substrates for oxidation when the direct supply of dietary nutrients has decreased and ATP requirements increase.}, number={9}, journal={JOURNAL OF NUTRITION}, author={Bruininx, Erik and Borne, Joost and Heugten, Eric and Milgen, Jaap and Verstegen, Martin and Gerrits, Walter}, year={2011}, month={Sep}, pages={1657–1663} } @article{fix_cassady_heugten_hanson_see_2010, title={Differences in lean growth performance of pigs sampled from 1980 and 2005 commercial swine fed 1980 and 2005 representative feeding programs}, volume={128}, ISSN={["1878-0490"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-76349085628&partnerID=MN8TOARS}, DOI={10.1016/j.livsci.2009.11.006}, abstractNote={The objective of this study was to assess how changes in genetics and feeding programs over 25 years in the U.S. commercial swine industry have impacted lean growth performance. Genetic samples (GS) of pigs (n = 162) from the commercial industries in 1980 and 2005 were randomly assigned to 1 of 2 feeding programs (FP) representative of 1980 or 2005. Pigs were placed 3 per pen (n = 54) at approximately 4 weeks of age and were harvested when average BW of the pen exceeded 116 kg. Real-time ultrasound measures for backfat depth and longissimus muscle area at the 10th rib were collected every 4 weeks, beginning at week 8 (group 1) or week 10 (group 2) until harvest. Average daily gain, ADFI, and G:F were calculated for the nursery period (7.0 ± 0.4 to 26.9 ± 0.7 kg BW), finishing period (26.9 ± 0.7 to 119 ± 0.7 kg BW), and overall (7.0 ± 0.4 to 116 ± 0.7 kg BW). Lean ADG and lean G:F were calculated for the period of first real-time ultrasound to harvest (42.7 ± 1.0 kg to 116 kg BW). Pigs from 2005 vs. 1980 GS and pigs fed 2005 vs. 1980 FP reached final BW of 116 kg sooner; 11 and 12 d, respectively. For ADG during finishing and overall, GS × FP interactions were observed, where 1980 GS pigs fed 1980 vs. 2005 FP showed increases of 7.0 and 6.3%, respectively; however, 2005 GS pigs fed 1980 vs. 2005 FP had increases of 12.6 and 12.3%, respectively. Pigs from the 2005 GS had greater ADG during finishing and overall, increased lean ADG, with no difference in ADFI during finishing, overall, and lean gain period or reduced ADFI during nursery which led to greater G:F and lean G:F. Pigs fed 2005 FP had increased ADG during all periods, with reduced ADFI during finishing, overall, and the lean gain test period which led to greater lean G:F and G:F during all portions of the trial. Although via different methods, changes over the past 25 years in the U.S. swine industry with respect to both genetics and feeding programs have resulted in a 15% reduction in days to harvest and a 45% improvement in lean efficiency.}, number={1-3}, journal={LIVESTOCK SCIENCE}, author={Fix, J. S. and Cassady, J. P. and Heugten, E. and Hanson, D. J. and See, M. T.}, year={2010}, month={Mar}, pages={108–114} } @article{heugten_hanson_ange_see_2010, title={Effects of on-farm magnesium supplementation through water on pork quality under two slaughter conditions}, volume={21}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-77954018450&partnerID=MN8TOARS}, DOI={10.1111/j.1745-4573.2009.00187.x}, abstractNote={ABSTRACT The objective of this study was to determine the effects of pre‐harvest Mg supplementation and stunning method on longissimus (LM) and semitendinosus (SM) pork muscle quality. Magnesium was supplemented for 2 days in drinking water (300 ppm) before pigs were harvested at a commercial packing plant using either electrical or CO2 stunning. Magnesium supplementation did not affect (P > 0.10) ultimate pH, initial fluid loss, retail display purge loss or color of the LM and SM. Pork quality was generally improved in LM and SM samples obtained from the plant using CO2 stunning as evidenced by reduced initial fluid loss (P < 0.001), reduced purge loss (P < 0.05) and lower L* (darker), a* (less red) and b* (less yellow) values (P < 0.05) during an 8‐day retail display. Pork quality was improved when using CO2 stunning technology; however, Mg supplementation did not improve pork quality in either plant.}, number={2}, journal={Journal of Muscle Foods}, author={Heugten, Eric and Hanson, D. and Ange, D. and See, M. T.}, year={2010}, pages={350–364} } @article{seabolt_van heugten_kim_ange-van heugten_roura_2010, title={Feed preferences and performance of nursery pigs fed diets containing various inclusion amounts and qualities of distillers coproducts and flavor1}, volume={88}, ISSN={0021-8812 1525-3163}, url={http://dx.doi.org/10.2527/jas.2009-2640}, DOI={10.2527/jas.2009-2640}, abstractNote={We evaluated the preferences of nursery pigs for diets containing increasing distillers dried grains with solubles (DDGS), varying in color, or high-protein distillers dried grains (HP-DDG) and the effects of flavor supplementation on pig preference and growth performance. In Exp. 1 through 5, diet preference was determined in weanling pigs adjusted to a commercial diet for at least 10 d, and then housed individually for a 2-d double-choice preference test. In Exp. 1, a total of 60 pigs (11.6 ± 0.3 kg of BW) were given a choice between a reference diet (0% DDGS) and test diets containing 0, 10, 20, or 30% DDGS. In Exp. 2, a total of 80 pigs (10.8 ± 0.1 kg of BW) were given a choice between a reference diet (0% HP-DDG) and diets containing 0, 10, 20, or 30% HP-DDG. In Exp. 3, a total of 80 pigs (10.3 ± 0.2 kg of BW) were given a choice between a reference diet (0% DDGS) and a diet containing 0%, 30% light, or 30% dark DDGS. In Exp. 4, a total of 80 pigs (11.2 ± 0.2 kg of BW) were given a choice between a reference diet without DDGS and a diet containing either 0% DDGS, 10 or 20% light DDGS, or 10 or 20% dark DDGS. In Exp. 5, a total of 108 pigs (9.0 ± 0.2 kg of BW) were given a choice between a reference diet (0% DDGS and no flavor) and a diet without or with flavor and containing 0, 10, or 20% DDGS. In Exp. 1 and 2, DDGS and HP-DDG, respectively, linearly decreased (P < 0.01) pig preference. In Exp. 3, dark DDGS were preferred (P < 0.05) compared with light DDGS. In Exp. 4, preferences were linearly reduced (P < 0.01) with DDGS inclusion, and dark DDGS tended (P = 0.06) to be preferred compared with light DDGS. In Exp. 5, DDGS reduced preference (P < 0.01) and flavor reduced preference (P < 0.01) regardless of DDGS level. In Exp. 6, a total of 192 pigs (6.7 ± 0.1 kg of BW) were fed starter 1 diets without or with flavor for 1 wk. Subsequently, pigs were fed starter 2 and 3 diets (2 wk each) containing 0, 10, or 20% DDGS while continuing to receive their respective flavor treatment. Flavor addition during the starter 1 phase increased ADFI (P = 0.02), and DDGS inclusion tended to decrease ADG (P = 0.06) and decreased ADFI (P = 0.03) during the starter 2 phase. Volatile components in DDGS and HP-DDG varied greatly depending on the source. Nursery pigs preferred a diet without DDGS or HP-DDG, and this appeared to be unrelated to color differences between sources. Knowledge of volatile compounds that enhance or suppress the palatability of feed may lead to further development of feed additives for masking relatively unpalatable, albeit cost-effective, ingredients.}, number={11}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Seabolt, B. S. and van Heugten, E. and Kim, S. W. and Ange-van Heugten, K. D. and Roura, E.}, year={2010}, month={Nov}, pages={3725–3738} } @article{kim_heugten_ji_lee_mateo_2010, title={Fermented soybean meal as a vegetable protein source for nursery pigs: I. Effects on growth performance of nursery pigs}, volume={88}, ISSN={["1525-3163"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-72649089143&partnerID=MN8TOARS}, DOI={10.2527/jas.2009-1993}, abstractNote={Four experiments were conducted using 671 nursery pigs to evaluate fermented soybean meal (FSBM) as a new vegetable protein source for nursery pigs. In Exp. 1, a total of 192 pigs weaned at 19.2 +/- 0.3 d of age were fed 3 diets (8 pens per treatment) for 2 wk: a control diet (without FSBM) and 2 diets with 3 and 6% FSBM replacing soybean meal, followed by a common diet for the next 2 wk. In Exp. 2, a total of 160 pigs weaned at 21.6 +/- 0.2 d of age were fed 4 diets (5 pens per treatment) for 2 wk: a control diet (without FSBM but with 25% dried skim milk; DSM) and 3 diets with 3, 6, and 9% FSBM replacing DSM based on CP. Concentrations of CP, Lys, Met, Thr, and Trp were kept consistent among diets in Exp. 1 and 2. In Exp. 3, a total of 144 pigs weaned at 22.1 +/- 0.2 d of age were fed 3 diets (6 pens per treatment) for 2 wk: a control diet (without FSBM but with 40% DSM) and 2 diets with 5 and 10% FSBM replacing DSM based on CP. Concentrations of CP, Lys, Met, Thr, Trp, and lactose were kept consistent among diets. In Exp. 4, a total of 175 pigs weaned at 20.7 +/- 0.4 d of age were fed 5 diets (5 pens per treatment) for 3 wk: a basal diet [15.5% CP without plasma protein (PP) and FSBM], 2 diets (18.4% CP) with 3.7% PP or 4.9% FSBM, and 2 diets (21.2% CP) with 7.3% PP or 9.8% FSBM. Concentrations of Lys, Met, Thr, and Trp were kept consistent among diets with the same CP concentrations. Pigs had access to feed and water ad libitum and their BW and feed intake were measured weekly for all experiments. Use of up to 6% FSBM replacing soybean meal improved (P < 0.05) G:F and diarrhea scores of nursery pigs (Exp. 1). Use of up to 9% FSBM replacing DSM reduced (P < 0.05) ADG and G:F (Exp. 2). When lactose concentrations were equal, FSBM could replace up to 10% DSM without adverse effects on ADG and G:F (Exp. 3). Relative bioavailability of protein in FSBM to PP was 99.1% (Exp. 4). Collectively, FSBM can serve as an alternative protein source for nursery pigs at 3 to 7 wk of age, possibly replacing the use of DSM and PP but excluding the first week postweaning for PP when balancing for AA and lactose.}, number={1}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Kim, S. W. and Heugten, E. and Ji, F. and Lee, C. H. and Mateo, R. D.}, year={2010}, month={Jan}, pages={214–224} } @article{sullivan_van heugten_ange-van heugten_poore_dierenfeld_wolfe_2009, title={Analysis of nutrient concentrations in the diet, serum, and urine of giraffe from surveyed North American zoological institutions}, volume={29}, ISSN={0733-3188 1098-2361}, url={http://dx.doi.org/10.1002/zoo.20278}, DOI={10.1002/zoo.20278}, abstractNote={Abstract}, number={4}, journal={Zoo Biology}, publisher={Wiley}, author={Sullivan, Kathleen and van Heugten, Eric and Ange-van Heugten, Kimberly and Poore, Matthew H. and Dierenfeld, Ellen S. and Wolfe, Barbara}, year={2009}, pages={n/a-n/a} } @article{ange-van heugten_van heugten_timmer_bosch_elias_whisnant_swarts_ferket_verstegen_2009, title={Fecal and Salivary Cortisol Concentrations in Woolly (Lagothrix ssp.) and Spider Monkeys (Ateles spp.)}, volume={2009}, ISSN={1687-8477 1687-8485}, url={http://dx.doi.org/10.1155/2009/127852}, DOI={10.1155/2009/127852}, abstractNote={Detrimental physiological effects due to stressors can contribute to the low captive success of primates. The objective of this research was to investigate the potential impact of diet composition on cortisol concentrations in feces and saliva in woolly (n=27) and spider monkeys (n=61). The research was conducted in three studies: the first investigated spider monkeys in the United States, the second investigated spider monkeys within Europe, and the third investigated woolly monkeys within Europe. Fecal cortisol in spider monkeys in US zoos varied (P=.07) from 30 to 66 ng/g. The zoo with the highest fecal cortisol also had the highest salivary cortisol (P≤.05). For European zoos, fecal cortisol differed between zoos for both spider and woolly monkeys (P≤.05). Spider monkeys had higher fecal cortisol than woolly monkeys (P≤.05). Zoos with the highest dietary carbohydrates, sugars, glucose, and fruit had the highest cortisol. Cortisol was highest for zoos that did not meet crude protein requirements and fed the lowest percentage of complete feeds and crude fiber. Differences among zoos in housing and diets may increase animal stress. The lifespan and reproductive success of captive primates could improve if stressors are reduced and dietary nutrients optimized.}, journal={International Journal of Zoology}, publisher={Hindawi Limited}, author={Ange-van Heugten, Kimberly D. and van Heugten, Eric and Timmer, Saskia and Bosch, Guido and Elias, Abahor and Whisnant, Scott and Swarts, Hans J. M. and Ferket, Peter and Verstegen, Martin W. A.}, year={2009}, pages={1–9} } @article{heugten_2009, title={Mycotoxins and Other Antinutritional Factors in Swine Feeds}, DOI={10.1201/9781420041842.ch25}, journal={Swine Nutrition, Second Edition}, publisher={CRC Press}, author={Heugten, Eric}, year={2009} } @article{ange-van heugten_verstegen_ferket_stoskopf_van heugten_2008, title={Serum chemistry concentrations of captive woolly monkeys (Lagothrix lagotricha)}, volume={27}, ISSN={0733-3188 1098-2361}, url={http://dx.doi.org/10.1002/zoo.20176}, DOI={10.1002/zoo.20176}, abstractNote={Abstract}, number={3}, journal={Zoo Biology}, publisher={Wiley}, author={Ange-van Heugten, Kimberly and Verstegen, Martin and Ferket, Peter R. and Stoskopf, Michael and van Heugten, Eric}, year={2008}, pages={188–199} } @article{ange-van heugten_burns_verstegen_jansen_ferket_heugten_2007, title={Evaluation of diabetes determinants in woolly monkeys (Lagothrix lagotricha)}, volume={91}, ISSN={["1439-0396"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-35948977775&partnerID=MN8TOARS}, DOI={10.1111/j.1439-0396.2007.00679.x}, abstractNote={Summary}, number={11-12}, journal={JOURNAL OF ANIMAL PHYSIOLOGY AND ANIMAL NUTRITION}, publisher={Wiley}, author={Ange-van Heugten, K. D. and Burns, R. and Verstegen, M. W. A. and Jansen, W. L. and Ferket, P. R. and Heugten, E.}, year={2007}, month={Dec}, pages={481–491} } @article{muley_heugten_moeser_rausch_kempen_2007, title={Nutritional value for swine of extruded corn and corn fractions obtained after dry milling}, volume={85}, ISSN={["1525-3163"]}, DOI={10.2527/jas.2006-127}, abstractNote={The experiment was designed to assess whether corn fractions or extrusion of corn can result in feed ingredients with a greater nutritional value than corn. Corn grain (8.0% CP, 0.21% P, 9.8% NDF) was processed by extrusion (82.8 degrees C, 345 kPa steam pressure for 12 s) or by dry milling to derive fractions rich in germ (13.1% CP, 1.19% P, 17.2% NDF), hulls (8.1% CP, 0.27% P, 32.6% NDF), and endosperm, namely tails (6.6% CP, 0.07% P, 3.6% NDF) and throughs (7.4% CP, 0.15% P, 4.5% NDF). Relative recovery in each fraction was 16, 20, 44, and 20%, respectively. Ileal digestibility of DM, P, and amino acids was determined using diets containing 7.0% CP from soybean meal and 5.3% CP from one of the test products. To allow for determination of standardized ingredient, ileal digestibility, basal endogenous AA losses were determined using a protein-free diet (74.6% cornstarch and 18.7% sucrose). Soybean meal ileal digestibility was determined using a diet (12.3% CP) based on soybean meal (23.3%). Eight barrows (27 +/- 2 kg) fitted with T-cannulas were fed 8 experimental diets (5-d adaptation and 2-d collection period) such that each diet was evaluated in at least 5 barrows. Relative to corn (77.9 +/- 1.2%), ileal digestibility of DM was greater for extruded corn (82.5%; P = 0.02), tails (85.9%; P < 0.01), and throughs (85.0%; P < 0.01), but it was lower for hulls (62.2%; P < 0.01) and germ (51.1%; P < 0.01). For P, corn (41.6 +/- 9.5%), throughs (47.2%), and hulls (57.3%) had similar ileal digestibility, but germ (7.9%) had lower ileal digestibility (P = 0.02) than corn; tails (27.6%) and extruded corn (23.5%) were not different from corn or germ but were lower than throughs and hulls. For total AA, corn (84.7 +/- 2.4%), throughs (84.3%), and hulls (85.8%) had similar ileal digestibility, but germ (76.6%) had lower ileal digestibility (P < 0.01) than corn; tails (82.0%) and extruded corn (81.7%) were intermediate. In conclusion, germ and hulls have a low ileal DM digestibility; germ also has low AA and P digestibility. Extrusion improved the ileal DM digestibility of corn. To maximize the ileal digestibility, removal of germ and hull from corn or extrusion of corn may thus be of interest.}, number={7}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Muley, N. S. and Heugten, E. and Moeser, A. J. and Rausch, K. D. and Kempen, T. A. T. G.}, year={2007}, month={Jul}, pages={1695–1701} } @article{eisemann_heugten_2007, title={Response of pigs to dietary inclusion of formic acid and ammonium formate}, volume={85}, ISSN={["0021-8812"]}, DOI={10.2527/jas.2006-464}, abstractNote={The objective of this study was to determine the optimal inclusion rate of dietary formic acid-ammonium formate (composition by weight was 62% formic acid and 37% ammonium formate) in nursery and grower-finisher diets or grower-finisher diets only. At weaning (d 21 +/- 2), 224 pigs (equal numbers of gilts and barrows) were blocked by BW within sex (28 pigs per BW block, 4 pigs per pen) and assigned randomly to 1 of 7 dietary treatments within each block. Dietary treatments (TRT), listed as percentage of dietary formic acid-ammonium formate in the nursery (NR) and the grower-finisher (GRF) diets, were as follows (NR and GRF): TRT 1: 0.0 and 0.0; TRT 2: 1.2 and 1.0; TRT 3: 0.0 and 1.0; TRT 4: 1.0 and 0.8; TRT 5: 0.0 and 0.8; TRT 6: 0.8 and 0.6; and TRT 7: 0.0 and 0.6. During the grower 2 (GR2) period, pigs fed treatments containing formic acid-ammonium formate in the nursery diets (TRT 2, TRT 4, and TRT 6) had greater (P < 0.05) ADG and G:F than pigs fed diets containing formic acid-ammonium formate in the grower period only (TRT 3, TRT 5, and TRT 7). Average daily feed intake tended to decrease (NR1, P = 0.07) or decreased (NR2, P < 0.05) for pigs fed formic acid-ammonium formate in the nursery (TRT 2, TRT 4, and TRT 6) compared with pigs fed control diets (TRT 1, TRT 3, TRT 5, and TRT 7). The ADFI also decreased (P < 0.05) during the GR1 and GR2 periods for pigs fed diets containing formic acid-ammonium formate compared with pigs fed control (TRT 1). In the combined nursery data, there was no effect (P > 0.10) of treatment on ADG. Pigs on diets containing formic acid-ammonium formate ate less feed (P < 0.05) and had improved G:F (P < 0.05) compared with pigs on the control treatments (TRT 1, TRT 3, TRT 5, and TRT 7). Combining the grower-finisher phases, G:F was greater (P = 0.05) for pigs fed diets containing formic acid-ammonium formate than for pigs fed the control feed. The efficiency of gain (i.e., G:F) was improved by 3.5% for pigs fed all formic acid-ammonium formate treatments and ranged from 2.3 (TRT 7) to 5.9% (TRT 4) compared with pigs fed control (TRT 1). Combining all phases from nursery to finisher, the G:F ratio tended (P = 0.08) to be greater for pigs fed formic acid-ammonium formate compared with pigs fed control. The efficiency of gain was improved by 3.0% for pigs fed all formic acid-ammonium formate treatments, ranging from 1.8 (TRT 7) to 5.2% (TRT 4), compared with pigs fed the control diet (TRT 1).}, number={6}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Eisemann, J. H. and Heugten, E.}, year={2007}, month={Jun}, pages={1530–1539} } @article{frederick_heugten_see_2006, title={Effects of pig age at market weight and magnesium supplementation through drinking water on pork quality}, volume={84}, ISSN={["1525-3163"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33744927165&partnerID=MN8TOARS}, DOI={10.2527/2006.8461512x}, abstractNote={Thirty-two halothane-negative pigs (109 +/- 0.6 kg of BW) were used to determine the effect of pig age at marketing (and thus growth rate), and magnesium supplementation through drinking water, on pork quality. Two initial groups of 50 pigs that differed by 30 +/- 2 d of age were fed diets to meet or exceed nutrient requirements beginning at 28 kg of BW. Sixteen average, representative pigs were selected from each group to represent older, slow-growing pigs and younger, fast-growing pigs. For the duration of the study, pigs were individually penned, provided 2.7 kg of feed (0.12% Mg) daily, and allowed free access to water. After 7 d of adjustment, pigs were blocked by sex and BW and allotted to 0 or 900 mg of supplemental Mg/L as MgSO4 in drinking water for 2 d before slaughter. All 32 pigs were then transported (110 km) to a commercial abattoir on the same day and slaughtered 2.5 h after arrival. Longissimus and semimembranosus (SM) chops were packaged and stored to simulate display storage for fluid loss and Minolta color determinations at 0, 2, 4, 6, and 8 d. Two remaining sections of the LM were vacuum-packaged and stored at 4 degrees C for 25 or 50 d. Fast- (younger) and slow- (older) growing pigs differed by 27 +/- 0.3 d of age (153 and 180 +/- 0.3 d; P < 0.001) at similar BW (108 and 110 +/- 0.6 kg of BW; P = 0.13). Supplementation of Mg tended to increase plasma Mg concentration (24.1 vs. 21.8 +/- 0.8 ppm; P = 0.06) but did not affect Mg concentration in LM or SM. Fluid loss of displayed LM or SM, and purge loss, color, and oxidation of vacuum-packaged LM or SM were not affected by age or Mg (P > 0.10). Surface exudate of the SM from older pigs was lower than that of younger pigs (61 vs. 74 +/- 6 mg; P = 0.05) but was not different for the LM (P = 0.22). The LM from older pigs displayed for 4 and 8 d; P < 0.05) were less yellow (lower b*) than younger pigs. The SM from older pigs had lower lightness (L*) initially (47.9 vs. 49.5 +/- 0.4) and after 2 d (49.7 vs. 51.1 +/- 0.4), 6 d (52.1 vs. 53.7 +/- 0.4) and 8 d (54.5 vs. 55.9 +/- 0.4) of display storage. Younger pigs had greater oxidation of the LM than older pigs on d 8 of display (P < 0.01), and Mg decreased oxidation on d 8 within younger pigs (P < 0.05). Pork quality was improved in older pigs as indicated by less exudate, reduced yellowness of the LM, reduced paleness of the SM, and reduced oxidation of the LM. However, Mg supplementation through the water for 2 d did not affect pork quality of either older, slower growing pigs or younger, faster growing pigs.}, number={6}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Frederick, B. R. and Heugten, E. and See, M. T.}, year={2006}, month={Jun}, pages={1512–1519} } @article{frederick_heugten_hanson_see_2006, title={Effects of supplemental magnesium concentration of drinking water on pork quality}, volume={84}, ISSN={["1525-3163"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-33744906659&partnerID=MN8TOARS}, DOI={10.2527/2006.841185x}, abstractNote={Thirty-two barrows were used to determine the effects of supplemental Mg in drinking water on pork quality. Pigs were determined to be free of the halothane and Napole mutations and were individually penned. After a 7-d adjustment period, barrows (111 +/- 1 kg BW) were blocked by BW and allotted randomly within block to 0, 300, 600, or 900 mg of supplemental Mg from Mg sulfate/L of drinking water for 2 d before slaughter. Pigs were not allowed access to feed (0.13% Mg) for 15 h before slaughter but continued to have access to experimental water treatments. Pigs were loaded and transported 110 km (1.75 h) to a commercial abattoir and remained in lairage for 5 h before slaughter. The LM was removed 24 h postmortem. Retail storage was simulated for 8 d, and the remaining LM was vacuum-packaged for 25 or 50 d at 4 degrees C. Plasma Mg concentration increased linearly (P = 0.001) with Mg supplementation; however, Mg concentration of the LM was not affected (P = 0.99) by Mg supplementation. Surface exudate, drip loss, and retail fluid loss of the LM were not affected (P > 0.10) by Mg. Lightness (L*) and redness (a*) of the LM were not affected (P > 0.10) by Mg, with the exception of initial redness (cubic; P = 0.05). Pigs supplemented with 300 or 900 mg of Mg/L had lower yellowness (b*) values of the LM displayed for 0 to 6 d than pigs supplemented with 0 or 600 mg of Mg/L (cubic; P < 0.05). Lightness of the LM after 25 (quadratic; P = 0.03) or 50 (quadratic; P = 0.04) d of vacuum-packed storage was greater at 300 and 600 mg of Mg/L than at 0 or 900 mg/L. Yellowness tended to be greater after 50 d, but not after 25 d, of vacuum-packaged storage for 300 or 600 mg of Mg/L compared with 0 or 900 mg/L (quadratic; P = 0.08). Oxidation of the LM, determined by thiobarbituric acid reactive substances after 4 d of retail storage, increased linearly (P = 0.05) as Mg increased in the drinking water. Furthermore, oxidation of the LM after 8 d of retail storage tended to increase linearly (P < 0.10), primarily because of the high oxidation of LM from pigs supplemented with 900 mg of Mg/L compared with controls (224 vs. 171 +/- 19 microg/kg, respectively). Oxidation of the LM was greater for pigs supplemented with 300 or 900 mg/L compared with 0 or 600 mg of Mg/L (cubic; P < 0.06) after 25 d of vacuum-packed storage. Magnesium did not improve pork quality characteristics of practical significance in pigs without the halothane and Rendement Napole mutations.}, number={1}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Frederick, BR and Heugten, E and Hanson, DJ and See, MT}, year={2006}, month={Jan}, pages={185–190} } @article{kempen_heugten_moeser_muley_sewalt_2006, title={Selecting soybean meal characteristics preferred for swine nutrition}, volume={84}, ISSN={["1525-3163"]}, DOI={10.2527/2006.8461387x}, abstractNote={As environmental constraints become more important issues for the animal industry, selecting feed ingredients that yield good animal performance but also minimize environmental impact of animal production becomes critical. The objective of this research was to identify which compositional features would be desirable for soybean meal to maximize nutritional value and minimize animal waste. Eight soybean samples were selected from a database of 72, such that maximal variability for CP, NDF, and ADF content was obtained. Samples were subsequently processed into meal using standardized procedures. In Experiment 1, 8 cannulated pigs were used to determine ileal digestibility following a Latin square design. In Experiment 2, 5 of the samples were used in complete feeds and 10 pigs were used in a crossover Latin square design to determine the total tract digestibility, odorants in fresh and 5-d-old manure, and ammonia emission from manure. Differences up to 6% in ileal DM digestibility and 8% in ileal CP digestibility were observed. This difference was reduced to 1.1% for total tract DM digestibility and 4% for total tract CP digestibility. Differences in odorant concentration were 3-fold and for in vitro ammonia emission were 42%. The only compositional variable with a significant effect on digestibility was stachyose, which negatively affected ileal digestibility of DM (r = -0.80, P = 0.02) and energy (r = -0.73, P = 0.04). None of the compositional variables measured affected ileal CP digestibility. Ileal CP digestibility, however, was correlated with estimated CP fermentation in the large intestine (r = -0.86, P = 0.06) and with in vitro ammonia emission after 48 h (r = -0.81, P = 0.09). In conclusion, nutritionally relevant variability exists in soy varieties. Low stachyose content is important for maximizing ileal energy digestibility of soybean meal. Although no compositional variable was identified that explained differences in ileal CP digestibility, maximizing ileal CP digestibility is of interest for maximizing the nutritional value of soybean meal and possibly for reducing ammonia and odor emissions.}, number={6}, journal={JOURNAL OF ANIMAL SCIENCE}, author={Kempen, T. A. T. G. and Heugten, E. and Moeser, A. J. and Muley, N. S. and Sewalt, V. J. H.}, year={2006}, month={Jun}, pages={1387–1395} } @article{comparison of dietary selenium fed to grower-finisher pigs from various regions of the united states on resulting tissue se and loin mineral concentration_2005, volume={83}, url={https://publons.com/wos-op/publon/19456802/}, DOI={10.2527/2005.834852x}, abstractNote={A study was conducted to evaluate the mineral content of pork tissue with particular emphasis on Se between various states (regions) having different diet (grain) indigenous Se concentrations. The study involved 19 states in the north, central, and southern regions of the United States, with committee members of NCR-42 and S-1012 (formerly S-288). A total of 62 pigs were used, with collaborators sending 100-g samples each of loin, heart, and liver, and a 3- to 4-g sample of hair (collected along the topline) from two to five market-weight pigs to a common laboratory for analysis. Diets at each station were formulated with locally purchased soybean meal and grain that was either grown or normally fed to pigs within their state. Tissues were analyzed for Se, but only the loin was analyzed for the macro- and micromineral elements. Correlation of dietary minerals to the tissue element was determined. The results demonstrated differences in tissue Se among states (P < 0.01), with high correlations of dietary Se to loin (r = 0.84; P < 0.01), heart (r = 0.84; P < 0.01), liver (r = 0.83; P < 0.01), and hair Se (r = 0.90; P < 0.01) concentrations. The correlation of hair Se to the Se concentration of loin, heart, and liver tissues was high (r > 0.90; P < 0.01). States in the west-central region of the United States and west of the Mississippi river had higher dietary Se and tissue Se concentrations than states in the eastern section of the Corn Belt, east of the Mississippi river, and along the East Coast. Generally, states did not differ greatly in their loin macro- and micromineral concentrations. The simple correlation of dietary minerals to their corresponding loin mineral concentration was generally non-significant, but most macrominerals had decreasing mineral concentrations when the dietary mineral level was higher. These results indicate that regional differences in tissue Se were influenced more by the indigenous Se content of the diet (grain) fed to the pigs than from sodium selenite.}, number={4}, journal={Journal of Animal Science}, year={2005}, pages={852–857} } @article{roberts_heugten_spears_routh_lloyd_almond_2004, title={Effects of dietary zinc on performance and immune response of growing pigs inoculated with porcine reproductive and respiratory syndrome virus and mycoplasma hyopneumoniae}, volume={17}, ISSN={["1976-5517"]}, DOI={10.5713/ajas.2004.1438}, abstractNote={The objective of this study was to determine the effects of dietary Zn level on performance, serum Zn concentrations, alkaline phosphatase activity (ALP), and immune response of pigs inoculated with Porcine Reproductive and Respiratory Syndrome virus (PRRSv) and Mycoplasma hyopneumoniae. A 2×4 factorial arrangement of treatments was used in a randomized design. Factors included; 1) PRRSv and M. hyopneumoniae inoculation (n=36 pigs) or sham inoculation (n=36 pigs) with media when pigs entered the grower facility (d 0) at 9 weeks of age and 2) 10, 50, 150 ppm supplemental Zn sulfate (ZnSO4) from weaning until the completion of the study, or 2,000 ppm supplemental ZnSO4 for two weeks in the nursery and then supplementation with 150 ppm ZnSO4 for the remainder of the trial. The basal diet contained 34 ppm Zn. Pigs were weighed on d 0, 10, 17, 24 and 31 and blood samples were collected on d 0, 7, 14, 21 and 28. Pigs inoculated with PRRSv were serologically positive at d 28 and control pigs remained negative to PRRSv. In contrast, the M hyopneumoniae inoculation was inconsistent with 33.3% and 52.8% of pigs serologically positive at d 28 in the control and infected groups, respectively. A febrile response was observed for approximately one week after inoculation with PRRSv. Feed intake (p<0.01) and gain (p<0.1) were less in PRRSv infected pigs than control pigs for the 31 d study. However, performance did not differ among pigs in the four levels of ZnSO4. Assessments of immune responses failed to provide unequivocal influence of either PRRSv inoculation or ZnSO4 level. These data suggest that PRRSv and M. hyopneumoniae act to produce some performance deficits and the influence of Zn supplementation of nursery age pigs does not have clear effect in grower pigs affected with disease. (Asian- Aust. J. Anim. Sci. 2004. Vol 17, No. 10 : 1438-1446)}, number={10}, journal={ASIAN-AUSTRALASIAN JOURNAL OF ANIMAL SCIENCES}, author={Roberts, ES and Heugten, E and Spears, JW and Routh, PA and Lloyd, KL and Almond, GW}, year={2004}, month={Oct}, pages={1438–1446} } @article{xing_heugten_li_touchette_coalson_odgaard_odle_2004, title={Effects of emulsification, fat encapsulation, and pelleting on weanling pig performance and nutrient digestibility}, volume={82}, DOI={10.2527/2004.8292601x}, abstractNote={Two experiments were conducted to evaluate the effect of lysolecithin on performance and nutrient digestibility of nursery pigs and to determine the effects of fat encapsulation by spray drying in diets fed in either meal or pelleted form. In Exp. 1, 108 pigs (21 d of age; 5.96 +/- 0.16 kg BW) were allotted to one of four dietary treatments (as-fed basis): 1) control with no added lard, 2) control with 5% added lard, 3) treatment 2 with 0.02% lysolecithin, and 4) treatment 2 with 0.1% lysolecithin in a 35-d experiment. Added lard decreased ADG (P = 0.02) and ADFI (P < 0.06) during d 15 to 35 and overall. Lysolecithin improved ADG linearly (P = 0.04) during d 15 to 35 and overall, but did not affect ADFI or G:F. Addition of lard decreased the digestibility of DM (P = 0.10) and CP (P = 0.05) and increased (P = 0.001) fat digestibility when measured on d 10. Lysolecithin at 0.02%, but not 0.10%, tended to improve the digestibility of fat (P = 0.10). On d 28, digestibilities of DM, fat, CP, P, (P = 0.001), and GE (P = 0.03) were increased with the addition of lard, and lysolecithin supplementation linearly decreased digestibilities of DM (P = 0.003), GE (P = 0.007), CP, and P (P = 0.001). In Exp. 2, 144 pigs (21 d of age, 6.04 +/- 0.16 kg BW) were allotted to one of six treatments in a 3 x 2 factorial randomized complete block design. Factors included 1) level (as-fed basis) and source of fat (control diet with 1% lard; control diet with 5% additional lard; and control diet with 5% additional lard from encapsulated, spray-dried fat) and 2) diet form (pelleted or meal). Addition of lard decreased feed intake during d 0 to 14 (P = 0.04), d 15 to 35 (P = 0.01), and overall (P = 0.008), and improved G:F for d 15 to 35 (P = 0.04) and overall (P = 0.07). Encapsulated, spray-dried lard increased ADG (P = 0.004) and G:F (P = 0.003) during d 15 to 28 compared with the equivalent amount of fat as unprocessed lard. Pelleting increased ADG (P = 0.006) during d 0 to 14, decreased feed intake during d 15 to 35 (P = 0.01), and overall (P = 0.07), and increased G:F during all periods (P < 0.02). Fat digestibility was increased (P = 0.001) with supplementation of lard, and this effect was greater when diets were fed in meal form (interaction, P = 0.004). Pelleting increased the digestibility of DM, OM, and fat (P < 0.002). Results indicate that growth performance may be improved by lysolecithin supplementation to diets with added lard and by encapsulation of lard through spray drying.}, number={9}, journal={Journal of Animal Science}, author={Xing, J. J. and Heugten, Eric and Li, D. F. and Touchette, K. J. and Coalson, J. A. and Odgaard, R. L. and Odle, Jack}, year={2004}, pages={2601–2609} } @article{pion_heugten_see_larick_pardue_2004, title={Effects of vitamin C supplementation on plasma ascorbic acid and oxalate concentrations and meat quality in swine}, volume={82}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-3142649749&partnerID=MN8TOARS}, DOI={10.2527/2004.8272004x}, abstractNote={Two experiments were conducted to determine the effects of vitamin C supplementation 48 h before slaughter on plasma ascorbic acid and oxalate concentrations and its effect on pork quality. In Exp. 1, 16 pigs (87.8+/-2.13 kg BW) were blocked by sex and weight and assigned randomly within block to one of three vitamin C treatments: 1) control; 2) 1,000 mg/L; or 3) 2,000 mg/L supplemented in the drinking water for a 48-h period. This was then followed by an additional 48-h period without supplemental vitamin C. Vitamin C increased plasma ascorbic acid concentrations (11.6, 19.5, and 23.4 microg/mL for 0, 1,000, and 2,000 mg/L of vitamin C; P < 0.05) within 6 h of supplementation. Plasma ascorbic acid concentrations from treated pigs decreased and did not differ from those of control pigs (13.7, 18.2, and 18.6 microg/mL for 0, 1,000, and 2,000 mg/L of vitamin C; P = 0.30) within 2 h of ending supplementation. No differences in plasma ascorbic acid concentrations were found between the two levels of supplementation. Vitamin C did not affect plasma oxalate or cortisol; however, cortisol tended to increase quadratically (P = 0.077) with vitamin C after 96 h. In Exp. 2, 30 pigs (107.5+/-0.54 kg BW) were blocked by sex and weight and assigned randomly within block to one of three vitamin C treatments: 1) control; 2) 500 mg/L; or 3) 1,000 mg/L supplemented in the drinking water 48 h before slaughter. Pigs were slaughtered 4 to 5 h after vitamin C supplementation ended, and loin samples were collected for meat quality measurements. At the time of slaughter, no differences in plasma ascorbic acid or cortisol were observed, but oxalate tended (P = 0.074) to increase quadratically with increasing vitamin C. Muscle ascorbic acid at slaughter and lactic acid in muscle at 0 and 1.5 h after slaughter were not different; however, lactic acid increased (P = 0.048) quadratically at 24 h after slaughter. Vitamin C did not affect initial or ultimate pH. Initial fluid loss (P = 0.041), and fluid loss on d 4 (P = 0.014) and 8 (P = 0.076) of simulated retail display; L* on d 0 (P = 0.038), 4 (P = 0.010), and 8 (P = 0.051); a* on d 0 (P = 0.021); and b* on d 0 (P = 0.006), 4 (P = 0.035), and 8 (P = 0.017) were negatively affected in a quadratic manner when vitamin C was supplemented. Vitamin C tended (P = 0.086) to increase oxidation in chops on d 0, but not d 4 or 8. Results indicate that on-farm supplementation of vitamin C was generally not effective in improving pork quality, which may be related to timing relative to slaughter.}, number={7}, journal={Journal of Animal Science}, author={Pion, S. J. and Heugten, Eric and See, M. T. and Larick, D. K. and Pardue, S.}, year={2004}, pages={2004–2012} } @article{van heugten_o'quinn_funderburke_flowers_spears_2004, title={Growth performance, carcass characteristics, plasma minerals, and fecal mineral excretion in grower-finisher swine fed diets with levels of trace minerals lower than common industry levels}, volume={12}, number={5}, journal={Journal of Swine Health and Production}, author={Van Heugten, E. and O'Quinn, P. R. and Funderburke, D. W. and Flowers, W. L. and Spears, J. W.}, year={2004}, pages={237–241} } @article{van heugten_frederick_2004, title={Magnesium supplementation and pork quality}, volume={25}, ISBN={0143-9014}, number={3}, journal={Pig News and Information}, author={Van Heugten, E. and Frederick, B. R.}, year={2004}, pages={101} } @article{frederick_heugten_see_2004, title={Timing of magnesium supplementation administered through drinking water to improve fresh and stored pork quality}, volume={82}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2442473252&partnerID=MN8TOARS}, DOI={10.2527/2004.8251454x}, abstractNote={Thirty-two pigs were used to determine the timing effect of magnesium (Mg) supplementation given through drinking water on pork quality. Pigs (16 barrows and 16 gilts) were individually penned, provided 2.7 kg of feed (0.12% Mg) daily (as-fed basis), and allowed free access to water via a nipple waterer for the duration of the study. After 5 d of adjustment, pigs (120 +/- 0.8 kg BW) were allotted randomly by weight and sex to 900 mg/L of supplemental Mg from magnesium sulfate heptahydrate in drinking water for -6, -4, -2, or 0 d relative to slaughter. The LM and semimembranosus (SM) muscles were removed 24 h postmortem. Retail display storage was simulated for 8 d, and the LM was vacuum-packaged for 25 or 50 d at 4 degrees C. Magnesium did not affect the pH of the LM at either 45 min (P = 0.15) or 24 h postmortem (P = 0.23). However, the pH of the SM at 24 h postmortem tended to be greater (P = 0.08) for pigs consuming Mg for 2 d than for those not supplemented. Fluid loss after 8 d of storage was less (P < 0.05) in the LM of pigs supplemented with Mg for 6 d than in those without supplementation. Furthermore, fluid loss from the SM of pigs provided supplemental Mg for 2 d, but not for 4 or 6 d, was lower (P < 0.05) on each day of retail display than the SM of unsupplemented pigs. Minolta L*, a*, and b* color measurements of the LM during display storage were not (P > 0.10) affected by Mg supplementation. However, Mg supplementation for 2 or 4 d decreased paleness (lower L* value) after 25 d (P < 0.05), but not 50 d (P > 0.10) of vacuum-packaged storage. Magnesium addition for 2 d decreased the extent of oxidation (thiobarbituric acid-reactive substances) of the LM after 4 d of display storage compared with 0 d of Mg (P < 0.05). Oxidation of the SM during 8 d of display storage increased linearly (P < 0.05) as duration of supplementation increased from 2 to 6 d but did not differ (P = 0.22) from 0 d of Mg supplementation. Although the response to Mg supplementation was variable, supplementation for 2 d before slaughter was considered most efficacious because of the following: decreased fluid loss from the SM, and lower lipid oxidation formation in the LM during retail storage; a darker, more desirable LM color after 25 d of vacuum-packaged storage; and cost reductions compared with longer durations.}, number={5}, journal={Journal of Animal Science}, author={Frederick, B. R. and Heugten, Eric and See, M. T.}, year={2004}, pages={1454–1460} } @article{moeser_see_heugten_morrow_kempen_2003, title={Diet and evaluators affect perception of swine waste odor: An educational demonstration}, volume={81}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-2142759654&partnerID=MN8TOARS}, DOI={10.2527/2003.81123211x}, abstractNote={An educational program was developed for extension agents, faculty, and graduate students to illustrate the effect of diet composition on odor from swine manure. Participants in this program first received a 2-h detailed review on odorous compounds in manure and the effect of diet on odor. For the second portion of the training, nine manure samples were used from pigs fed diets formulated with feed ingredients predicted to have different effects on odor emission or a nutritionally adequate corn-soybean meal diet. Participants were instructed to rate the odor from these samples for pleasantness, irritation, and intensity on a scale of 0 (best) to 8 (worst), using manure from the corn-soybean meal fed pig as the reference with a score defined as 4 for each variable. Results obtained were summarized and discussed before concluding the program. Participants were Cooperative Extension Agents (n = 13) with swine responsibilities and graduate students and faculty (n = 8). The manure from the diet with the worst odor scores (1% garlic) was rated at 70% more odorous across the three odor variables (P < 0.05) than the diet with the least odorous manure (purified diet). Even though a reference sample was used, individual participants differed in their perception of irritation across samples (P < 0.05), ranging in average score across diets from 2.4 (moderately better than reference) to 5.0 (slightly worse than reference). With extension agents, a 1 to 7 scale (very interesting to not at all interesting) was used for evaluation of the training session. Participants found the material to be interesting (mean = 1.7, SD = 0.7) and the training exercise to be well organized and coherent in its presentation (mean = 1.8, SD = 0.7). Participants enjoyed this training and learned that differences in odor are achievable through altering diet composition, and that the response to swine odor depends on individual odor perception.}, number={12}, journal={Journal of Animal Science}, author={Moeser, A. J. and See, M. T. and Heugten, Eric and Morrow, W. E. M. and Kempen, T. Van}, year={2003}, pages={3211–3215} } @article{heugten_spears_kegley_ward_qureshi_2003, title={Effects of organic forms of zinc on growth performance, tissue zinc distribution, and immune response of weanling pigs}, volume={81}, DOI={10.2527/2003.8182063x}, abstractNote={This study was conducted to determine the effect of zinc level and source on growth performance, tissue Zn concentrations, intracellular distribution of Zn, and immune response in weanling pigs. Ninety-six 3-wk-old crossbred weanling pigs (BW = 6.45 +/- 0.17 kg) were assigned to one of six dietary treatments (four pigs per pen, four replicates per treatment) based on weight and litter origin. Treatments consisted of the following: 1) a corn-soybean meal-whey diet (1.2% lysine) with a basal level of 80 ppm of supplemental Zn from ZnSO4 (control; contained 104 ppm total Zn); 2) control + 80 ppm added Zn from ZnSO4; 3) control + 80 ppm added Zn from Zn methionine (ZnMet); 4) control + 80 ppm added Zn from Zn lysine (ZnLys); 5) control + 40 ppm added Zn from ZnMet and 40 ppm added Zn from ZnLys (ZnML); and 6) control + 160 ppm added Zn from ZnSO4. Zinc supplementation of the control diet had no effect on ADG or ADFI. Gain efficiency was less (P < 0.05) for pigs fed 80 ppm of Zn from ZnSO4 than for control pigs and pigs fed 160 ppm of Zn from ZnSO4. Organ weights, Zn concentration, and intracellular distribution of Zn in the liver, pancreas, and spleen were not affected (P = 0.12) by Zn level or source. Skin thickness response to phytohemagglutinin (PHA) was not affected (P = 0.53) by dietary treatment. Lymphocyte proliferation in response to PHA was greater (P < 0.05) in pigs fed ZnLys than in pigs fed the control diet or the ZnML diet; however, when pokeweed mitogen was used, lymphocyte proliferation was greatest (P < 0.05) in pigs fed the ZnMet diet than pigs fed the control, ZnLys, ZnML, or 160 ppm ZnSO4 diets. Antibody response to sheep red blood cells was not affected by dietary treatments. Supplementation of 80 ppm of Zn from ZnSO4 or ZnMet and 160 ppm of Zn from ZnSO4 decreased (P < 0.05) the antibody response to ovalbumin on d 7 compared with control pigs, but not on d 14. Phagocytic capability of peritoneal exudate cells was increased (P < 0.05) when 160 ppm of Zn from ZnSO4 was supplemented to the diet. The number of red blood cells ingested per phagocytic cell was increased (P < 0.05) in pigs fed the diet supplemented with a combination of ZnMet and ZnLys and the diet with 160 ppm of Zn from ZnSO4. Results suggest that the level of Zn recommended by NRC for weanling pigs was sufficient for optimal growth performance and immune responses, although macrophage function may be enhanced at greater levels of Zn. Source of Zn did not alter these measurements.}, number={8}, journal={Journal of Animal Science}, author={Heugten, Eric and Spears, J. W. and Kegley, E. B. and Ward, J. D. and Qureshi, M. A.}, year={2003}, pages={2063–2071} } @article{heugten_funderburke_dorton_2003, title={Growth performance, nutrient digestibility, and fecal microflora in weanling pigs fed live yeast}, volume={81}, DOI={10.2527/2003.8141004x}, abstractNote={Two experiments were conducted to evaluate the effects of live yeast supplementation on nursery pig performance, nutrient digestibility, and fecal microflora and to determine whether live yeast could replace antibiotics and growth-promoting concentrations of Zn and Cu in nursery pigs. In Exp. 1, 156 pigs were weaned at 17 d of age (BW = 5.9 kg) and allotted to a 2 x 2 factorial randomized complete block design (six or seven pigs per pen with six pens per treatment). Factors consisted of 1) dietary supplementation with oat products (oat flour and steam-rolled oats; 0 or 27.7%) and 2) yeast supplementation at 0 or 1.6 x 10(7) cfu of Saccharomyces cerevisiae SC47/g of feed. In Exp. 2, 96 pigs were weaned at 17 d of age and allotted to a 2 x 2 factorial randomized complete block design (four pigs per pen with six pens per treatment) with factors of 1) diet type (positive control containing growth-promoting concentrations of Zn, Cu, and antibiotics or negative control) and 2) live yeast supplementation (0 or 2.4 x 10(7) cfu of Saccharomyces cerevisiae SC47/g of feed). The inclusion of oat products in Exp. 1 decreased (P < 0.10) overall ADG and final BW. Yeast supplementation did not affect growth performance of pigs in Exp. 1 (P = 0.65); however, ADG in Exp. 2 was 10.6% greater (P < 0.01) and ADFI was increased by 9.4% (P < 0.10) in pigs supplemented with yeast in the positive control diet. Addition of Zn, Cu, and antibiotics to the diet improved gain:feed ratio during the prestarter period (P < 0.02) and overall (P = 0.10). In Exp. 1, inclusion of oat products increased (P < 0.01) total bacteria in feces when measured on d 10. Fecal lactobacilli measured on d 28 were reduced (P < 0.05) in pigs fed diets with oat products and yeast (interaction, P < 0.05). In Exp. 2, yeast supplementation decreased (P < 0.05) total bacteria and lactobacilli. Dietary yeast resulted in a greater (P < 0.05) yeast count in feces of pigs during the starter phase of Exp. 1. Yeast decreased (P < 0.10) the digestibility of DM, fat, and GE in the prestarter phase and DM, fat, P, and GE in the starter phase, whereas oat products increased the digestibility of DM, CP, fat, and GE (P < 0.05) in the prestarter phase. Results indicate that live yeast supplementation had a positive effect on nursery pig performance when diets contained growth-promoting antimicrobials. Nonetheless, the response was variable, and the conditions under which a response might be expected need to be further defined.}, number={4}, journal={Journal of Animal Science}, author={Heugten, Eric and Funderburke, D. W. and Dorton, K. L.}, year={2003}, pages={1004–1012} } @article{kempen_baker_heugten_2003, title={Nitrogen losses in metabolism trials}, volume={81}, DOI={10.2527/2003.81102649x}, abstractNote={The utilization of dietary nitrogen has been the subject of much research. For example, NRC (1998) provides a wealth of information on the subject that is based on an extensive review of literature. For a 45-kg pig fed a corn-soybean meal diet, the NRC predicts an efficiency of N utilization of 35%. This low efficiency is the result of approximately 7% of the dietary CP not being digestible and the equivalent of 8% of the dietary CP being lost in endogenous excretions. The synthesis of endogenous material results in the obligatory catabolism of the equivalent of 10% of the dietary N. These losses account for 25% of the dietary N intake. The remaining CP (75% of dietary) is available for lean tissue accretion; however, as a result of mismatches between requirements and dietary supply, approximately 30% is degraded and used for energy production. The remainder is actually used for lean tissue growth, but approximately 10% is not accreted due to inefficiencies in lean tissue growth, and summation of these losses results in N utilization in the neighborhood of 35% (van Kempen and van Heugten, 2000). Feed wastage and poor animal health can deteriorate this efficiency even further. Chung and Baker (1992) demonstrated that with diets formulated to be nearly 100% digestible (which would minimize indigestible and endogenous losses and optimally match the requirement of the animal, thereby minimizing the 30% mismatch), efficiencies of 60% were achievable in nursery pigs, lending credibility to the above calculation.}, number={10}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Kempen, T. A. T. G. Van and Baker, D. H. and Heugten, Eric}, year={2003}, pages={2649–2650} } @article{heugten_hasty_see_larick_2003, title={Storage stability of pork from Berkshire and Hampshire sired pigs following dietary supplementation with vitamin E}, volume={14}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0037353707&partnerID=MN8TOARS}, DOI={10.1111/j.1745-4573.2003.tb00346.x}, abstractNote={ABSTRACT}, number={1}, journal={Journal of Muscle Foods}, author={Heugten, Eric and Hasty, J. L. and See, M. T. and Larick, D. K.}, year={2003}, pages={67–80} } @article{roberts_heugten_lloyd_almond_spears_2002, title={Dietary zinc effects on growth performance and immune response of endotoxemic growing pigs}, volume={15}, ISSN={["1976-5517"]}, DOI={10.5713/ajas.2002.1496}, abstractNote={A 2×3 factorial arrangement of treatments was used in a completely randomized design to determine the effects of dietary Zn on performance and immune response of acutely endotoxemic growing pigs (n=96, mean BW=24.9 kg). Factors included 1) intramuscular injection of 10 µg/kg BW of Escherichia coli lipopolysaccharide (LPS) or control and 2) supplemental Zn at 10, 50, or 150 ppm. Diets were fed beginning after weaning (initial body weight=7.6 kg) in the nursery and continued for 16 d into the grower phase. The basal corn-soybean meal grower diet contained 1% lysine and 34.3 ppm Zn. Pigs were acclimated for 12 d in the grower- finishing facility before LPS treatment on d 13. Gain, feed intake, and feed efficiency were unaffected by dietary Zn. Feed intake decreased (p<0.10) and gain/feed was greater (p<0.10) from d 13 to d 16 for pigs injected with LPS. Serum Zn and alkaline phosphatase activity increased (p<0.05) with increasing Zn levels. The febrile response to LPS peaked at 6 h post exposure and pigs were afebrile within 12 h. Rectal temperature was greater (p<0.05) in pigs receiving 50 and 150 ppm Zn than in pigs supplemented with 10 ppm Zn. In vivo cellular immune response, measured on d 13 by skin thickness response to phytohemagglutinin (PHA), was greater after 6 h (p< 0.05) in pigs fed 10 ppm Zn and exposed to LPS compared to all other treatments, but was not affected at 12, 24 or 48 h. Zinc did not affect mitogen induced lymphocyte proliferation. Zinc supplemented at 50 or 150 ppm resulted in an enhanced febrile response in pigs subjected to iatrogenic endotoxemia, but did not affect pig performance or immune response measurements. (Asian-Aust. J. Anim. Sci. 2002. Vol 15, No. 10 : 1496-1501)}, number={10}, journal={ASIAN-AUSTRALASIAN JOURNAL OF ANIMAL SCIENCES}, author={Roberts, ES and Heugten, E and Lloyd, K and Almond, GW and Spears, JW}, year={2002}, month={Oct}, pages={1496–1501} } @article{hasty_heugten_see_larick_2002, title={Effect of vitamin E on improving fresh pork quality in Berkshire- and Hampshire-sired pigs}, volume={80}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0038179609&partnerID=MN8TOARS}, DOI={10.2527/2002.80123230x}, abstractNote={This study was designed to evaluate the effects of vitamin E supplementation on pork quality of two genotypes with distinct differences in pork quality traits. Pigs (n = 240; BW = 87 +/- 0.35 kg) were allotted by weight to one of 20 treatments (4 pens/treatment, 3 pigs/pen) in a 2 x 2 x 5 factorial randomized complete block design. Factors included 1) genotype (Berkshire or Hampshire sired), 2) sex (gilts or barrows), and 3) vitamin E level (12.1, 54.7, 98.8, 174.0, and 350.6 IU of vitamin E/kg diet). Hampshire-sired pigs had greater average daily gain (1.05 vs 0.98 kg) and gain:feed (0.30 vs 0.27) and less average daily feed intake (ADFI) (3.46 vs 3.62 kg) than Berkshire-sired pigs (P < 0.001) for the 6-wk study. Hampshire-sired barrows consumed more feed (3.54 vs 3.38 kg/d) and were less efficient (0.29 vs 0.31) than Hampshire-sired gilts (P < 0.05), but this sex difference was not observed in Berkshire-sired pigs (interaction, P < 0.05). Berkshire-sired pigs had greater backfat (34.1 vs 21.1 mm; P < 0.001), reduced longissimus muscle area (37.6 vs 46.3 cm2; P < 0.001), reduced lean percentage (53.0 vs 55.8; P < 0.001), and a greater head-on yield (79.8 vs 79.2; P < 0.05). Vitamin E increased (P < 0.05) ADFI linearly (P < 0.05), but had no effects on carcass composition. Loin chops from Hampshire-sired pigs had reduced ultimate pH (5.64 vs 5.91), greater drip loss (92.2 vs 66.3 mg), and increased Minolta L* (52.6 vs 48.6), a* (8.9 vs 7.5), and b* (6.9 vs 5.2) values compared to Berkshire-sired pigs (P < 0.001). Vitamin E had no effect on pH, temperature, drip loss, and L* or a* values, but tended (P < 0.07) to increase b* values linearly (P < 0.06). Oxidation as indicated by thiobarbituric acid reactive substances (TBARS) was greatest in Hampshire-sired gilts at the lowest level of vitamin E, and decreased linearly (P < 0.001) with additional vitamin E. However, TBARS responded in a cubic fashion (P < 0.05) to vitamin E in Hampshire-sired barrows and were not affected in Berkshire-sired gilts or barrows (three-way interaction, P < 0.02). Hampshire-sired pigs had greater TBARS than Berkshire-sired pigs (0.053 vs 0.047 mg malondialdehyde equivalents/kg). Vitamin E supplementation increased serum concentrations of vitamin E on d 21 (1.06 to 4.79 microg/mL) and d 42 (1.02 to 2.82 microg/mL) and increased tissue concentrations of vitamin E (1.99 to 4.83 microg/g) linearly (P < 0.001). Vitamin E supplementation was not effective in improving fresh meat quality in genotypes with poor or superior meat quality traits.}, number={12}, journal={Journal of Animal Science}, author={Hasty, J. L. and Heugten, Eric and See, M. T. and Larick, D. K.}, year={2002}, pages={3230–3237} } @article{heugten_kempen_2002, title={Growth performance, carcass characteristics, nutrient digestibility and fecal odorous compounds in growing-finishing pigs fed diets containing hydrolyzed feather meal}, volume={80}, DOI={10.2527/2002.801171x}, abstractNote={This study was designed to determine the effects of hydrolyzed feather meal inclusion on growth performance, carcass characteristics, nutrient digestibility and fecal odorous compounds in modern lean growth genotype pigs. Two hundred forty pigs (BW = 23.2 +/- 1.3 kg) were allotted based on BW and sex to a 2 x 6 factorial arrangement of treatments (four pens per treatment; five pigs per pen) in a randomized complete block design. Factors consisted of 1) sex (barrows or gilts) and 2) dietary treatment (0, 2, 4, 6, 8, or 10% hydrolyzed feather meal). Diets were formulated to contain 1.00, 0.90, 0.75, or 0.60% apparent ileal digestible lysine for phases 1 to 4, respectively, with other amino acids provided at an ideal ratio. Available P and ME were kept constant within each phase. No significant interactions between feather meal inclusion and sex were observed for growth performance (P > 0.15). Body weight gain was reduced (P < 0.05) for pigs fed 10% feather meal compared to pigs fed 0, 4, or 8% feather meal. Feed intake of pigs fed 10% feather meal was reduced (P < 0.05) compared to pigs fed 0 or 4% feather meal. Ultrasound backfat measurements tended (P = 0.12) to increase with increasing levels of feather meal. Daily lean gain was less (P < 0.05) in pigs fed 10% feather meal than in pigs fed either 0, 2, 4, or 8% feather meal. Digestibility of N measured on wk 9 decreased quadratically (P < 0.001) with increasing levels of feather meal. Phosphorus digestibility increased in a linear fashion (P < 0.02), however, the improvement in P digestibility with increasing levels of feather meal was more pronounced in barrows compared to gilts (interaction, P < 0.05). Fecal samples obtained from pigs fed 0, 4, or 8% feather meal were analyzed for odorous compounds. Concentrations of butanoic, pentanoic, and 3-methylbutanoic acid were greater (P < 0.05) and concentrations of 3-methylphenol, 4-methylphenol, indole, and decane were less (P < 0.05) in feces from pigs fed feather meal. These results suggest that feather meal can be included in diets for growing-finishing pigs at a rate of 8%. Excretion of N in feces increased but P excretion decreased with increasing levels of feather meal. Odorous compounds in feces can be affected by the inclusion of hydrolyzed feather meal, but the exact impact of these changes on odor perception remains to be elucidated.}, number={1}, journal={Journal of Animal Science}, author={Heugten, Eric and Kempen, T. Van}, year={2002}, pages={171–178} } @article{ferket_heugten_kempen_angel_2002, title={Nutritional strategies to reduce environmental emissions from nonruminants}, volume={80}, DOI={10.2527/animalsci2002.80e-suppl_2e168x}, abstractNote={The amount of nutrients (i.e., N, P, Zn, and Cu) and associated odors emitted from production animals into the environment can be modulated by sev- eral different nutritional strategies, but their practical application is dependent on costs and biological limita- tions. In general, nutrient excretion may be reduced by avoiding the overfeeding of specific nutrients or by us- ing nutritional manipulations to enhance nutrient utili- zation in the animal. Loss to the environment can be avoided by manufacturing and handling the feed in a pelletized form that will minimize waste and improve feed/gain. Other strategies for minimizing nutrient losses include: 1) the development of feeding programs that are specific for sex and strain of the animal; 2) increasing the number of feed phases to better meet the animal's age-related requirements; 3) formulating diets to include the minimal amounts of nutrients re- quired to satisfy production goals; 4) meeting the ani- mal's amino acid requirements; 5) using synthetic amino acid supplements to feed to reduce N emission; 6) using feed ingredients with high digestibility and nutrient bioavailability; and 7) formulating diets based on nutrient availability instead of total nutrient con- tent. Nutrient digestibility of feedstuffs is dependent on processing conditions, genetic characteristics of the grains and oilseeds, and the presence of nutritional antagonists in specific feedstuffs used in the diet. Feed ingredients that lead to odor production can be avoided (e.g., fishmeal and some easily fermentable feed ingre- dients). Feed additives, such as antibiotics, nonstarch polysaccharides, direct-fed microbials, organic acids, microbial enzymes (i.e., phytase, carbohydrases, and proteases) can be used to increase the digestibility and absorption of nutrients or to modulate the microflora. Finally, a cost factor for the control or disposal of nutri- ents or odor should be considered in the feed formula- tion to optimize the various nutritional strategies dis- cussed above. Regardless of biological and economic limitations, significant reductions in nutrient and odor emission from nonruminants can be achieved by appro- priate nutritional strategies, but response may differ for swine and poultry.}, number={E-suppl_2}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Ferket, P. R. and Heugten, E. and Kempen, T. A. T. G. and Angel, R.}, year={2002}, month={Jan}, pages={E168–E182} } @article{kempen_kim_heugten_2002, title={Pigs as recyclers for nutrients contained in Bermuda grass harvested from spray fields}, volume={81}, ISSN={["0960-8524"]}, DOI={10.1016/S0960-8524(01)00129-8}, abstractNote={The ability of pigs to use nitrogen and energy in Bermuda grass was evaluated in order to assess whether Bermuda grass harvested from spray fields could be fed to pigs as a means to recycle nitrogen. Digestibility of Bermuda grass incorporated into corn-soybean meal diets was evaluated in heavy finishing pigs and gestating sows. Results suggest that Bermuda grass digestibility is negative in animals not adapted to a high-fiber diet. Enzymes improve this digestibility, but even with enzymes, nitrogen digestibility was poor. Pigs fed a diet containing 10% Bermuda grass required a one week adaptation period for maximal digestion; following adaptation, pigs can digest approximately 40% of the energy in Bermuda grass but none of the nitrogen. Feeding Bermuda grass to pigs as a means of recycling nitrogen is thus not recommended.}, number={3}, journal={BIORESOURCE TECHNOLOGY}, publisher={Elsevier BV}, author={Kempen, TATG and Kim, I and Heugten, E}, year={2002}, month={Feb}, pages={233–239} } @article{moeser_kim_heugten_kempen_2002, title={The nutritional value of degermed, dehulled corn for pigs and its impact on the gastrointestinal tract and nutrient excretion}, volume={80}, DOI={10.2527/2002.80102629x}, abstractNote={Three experiments were designed to assess the feeding value and potential environmental benefits of feeding degermed, dehulled corn, a low fiber by-product originating from the corn dry milling process, to pigs. Twelve 27-kg (SE = 0.8) barrows were used in Exp. 1 to measure the apparent fecal digestibility of DM, GE and N of degermed, dehulled corn compared with corn grain. Two diets were formulated to contain either 96.4% of degermed, dehulled corn or corn grain plus supplemental vitamins and minerals. Digestibilities of DM, GE, and N were greater in degermed, dehulled corn (96.2, 96.0, and 93.6%, respectively) compared with corn grain (89.0, 89.0, and 78.4%, respectively) (P < 0.01). Overall, a 67 and 29% reduction in DM and N excretion, respectively, was observed. In Exp. 2, eight 70-kg (SE =1.8) barrows were surgically fitted with ileal cannulae and fed the same diets as in Exp. 1, to measure the ileal digestibility of nutrients in degermed, dehulled corn. Ileal digestibility of DM, energy, and N was 13, 15, and 7% greater in degermed, dehulled corn (P < 0.05). Apparent ileal digestibility coefficients of leucine, methionine, and phenylalanine were greater in degermed, dehulled corn compared with corn grain (P < 0.05) while a trend for a lower tryptophan digestibility in degermed, dehulled corn was observed (P = 0.067). In Experiment 3, 96 nursery pigs with an initial average BW of 8.8 kg (SE = 0.08), fed a starter diet formulated with degermed, dehulled corn or corn grain as the major grain source, were used in a 28-d growth performance study. At the end of the study, 24 pigs (1 pig per pen) were sacrificed and gastrointestinal tract measurements were taken. Daily growth rates of pigs were the same between diets (0.64 kg/d). A trend for reduced feed intake (P = 0.073) in pigs fed degermed, dehulled corn led to a 4% improvement in gain to feed (P < 0.05). Feeding degermed, dehulled corn had no effect on gut fill, gastrointestinal tract weight, or liver weight (P > 0.05). Ileal villus lengths and crypt depths were not affected by feeding degermed, dehulled corn although ileal villus widths were greater in pigs fed corn grain. Results from these trials suggest that corn processed to remove poorly digestible fiber fractions provides more digestible nutrients than corn grain. As a result, degermed, dehulled corn reduces fecal and N excretion, thus providing a means to reduce nutrient excretion.}, number={10}, journal={Journal of Animal Science}, author={Moeser, A. J. and Kim, I. B. and Heugten, Eric and Kempen, T. Van}, year={2002}, pages={2629–2638} } @article{kempen_heugten_trottier_2001, title={Adipic acid increases plasma lysine but does not improve the efficiency of lysine utilization in swine}, volume={79}, DOI={10.2527/2001.7992406x}, abstractNote={Adipic acid, upon catabolism, results in intermediates that bear a structural similarity to lysine degradation products. The objectives of this research were to determine whether adipic acid affects lysine concentrations in plasma and to evaluate whether adipic acid improves the efficiency of lysine utilization in pigs. In Exp. 1, nursery pigs (n = 14) were fed (for a period of 7 d) either a standard nursery diet or the same diet supplemented with 1% adipic acid to assess effects on plasma amino acid concentrations (plasma collected on d 7). In Exp. 2, nursery pigs (n = 56) were fed (for a period of 15 d) either a control diet or the same diet but deficient in either lysine, threonine, or tryptophan with or without supplemental adipic acid to assess the effects of adipic acid on the efficiency of amino acid utilization. The results from Exp. 1 showed that adipic acid increased plasma lysine (by 18%) but not alpha-amino adipic acid, an intermediate in lysine degradation. Experiment 2 demonstrated that adipic acid did not increase the efficiency of utilization of lysine, threonine, or tryptophan. The lack of effects on alpha-amino adipic acid in Exp. 1 and the lack of a positive effect on the efficiency of utilization of lysine, threonine, and tryptophan suggest that adipic acid does not inhibit the mitochondrial uptake of lysine and(or) its degradation in the mitochondrion. It is concluded that feeding adipic acid increases plasma lysine but does not improve the efficiency of lysine utilization.}, number={9}, journal={Journal of Animal Science}, publisher={Oxford University Press (OUP)}, author={Kempen, T. Van and Heugten, Eric and Trottier, N. L.}, year={2001}, pages={2406–2411} } @article{heo_odle_han_cho_seo_heugten_pilkington_2000, title={Dietary L-carnitine improves nitrogen utilization in growing pigs fed low energy, fat-containing diets}, volume={130}, ISSN={["0022-3166"]}, DOI={10.1093/jn/130.7.1809}, abstractNote={Growing pigs (n = 25; 17.8 +/- 0.1 kg) were used to study the effects of L-carnitine and protein intake on nitrogen (N) balance and body composition. Fat-supplemented (40 g soy oil/kg diet), corn-soybean meal basal diets containing low or high protein (136 or 180 g/diet) were formulated so that protein accretion would be limited by metabolizable energy (ME). Each basal diet was supplemented with 0 or 500 mg/kg L-carnitine and fed to pigs for 10 d in a nutrient balance trial. Final body composition was compared with weight and age-matched pigs measured on d 0 to calculate nutrient accretion rates. High protein feeding increased (P < 0.01) average daily gain (ADG) by 34%, as well as nitrogen digestibility (4.4%), retention (5.2%), urinary excretion (29%) and crude protein (CP) accretion (33%). Total-body carnitine accretion rate was 4.5 fold greater and total body carnitine concentration was almost 100% greater than in unsupplemented controls (P < 0.01). Irrespective of protein level, carnitine increased ADG (by 7.3%, P < 0.10) and CP accretion rate (9%, P < 0.10). Congruently, carnitine supplementation improved the efficiency of nitrogen retention (P < 0. 05) and reduced urinary nitrogen excretion (14%, P < 0.10). Carcass fat content also was reduced in carnitine-supplemented pigs (P < 0. 10). Collectively, these data support the hypothesis that carnitine can improve the efficiency of nitrogen utilization in 20-kg pigs fed energy-limited, fat-containing diets. We conclude that endogenous carnitine biosynthesis may be adequate to maintain sufficient tissue levels during growth, but that supplemental dietary carnitine (at 500 mg/kg) may be retained sufficiently so as to alter nutrient partitioning and thus body composition of 20-kg pigs.}, number={7}, journal={JOURNAL OF NUTRITION}, author={Heo, K and Odle, J and Han, IK and Cho, W and Seo, S and Heugten, E and Pilkington, DH}, year={2000}, month={Jul}, pages={1809–1814} } @article{armstrong_spears_heugten_engle_wright_2000, title={Effect of copper source (cupric citrate vs cupric sulfate) and level on growth performance and copper metabolism in pigs}, volume={13}, ISSN={["1011-2367"]}, DOI={10.5713/ajas.2000.1154}, number={8}, journal={ASIAN-AUSTRALASIAN JOURNAL OF ANIMAL SCIENCES}, author={Armstrong, TA and Spears, JW and Heugten, E and Engle, TE and Wright, CL}, year={2000}, month={Aug}, pages={1154–1161} } @article{van heugten_van kempen_1999, title={Methods may exist to reduce nutrient excretion}, volume={71}, number={17}, journal={Feedstuffs}, author={Van Heugten, E. and Van Kempen, T.}, year={1999}, pages={12–13} } @article{van heugten_sweet_stumpf_risley_schell_1997, title={Effects of water supplementation with selenium and vitamin E on growth performance and blood selenium and serum vitamin E concentrations in weanling pigs}, volume={211}, number={8}, journal={Journal of the American Veterinary Medical Association}, author={Van Heugten, E. and Sweet, L. A. and Stumpf, T. T. and Risley, C. R and Schell, T. C.}, year={1997}, pages={1039-} } @article{heugten_spears_1997, title={Immune response and growth of stressed weanling pigs fed diets supplemented with organic or inorganic forms of chromium}, volume={75}, DOI={10.2527/1997.752409x}, abstractNote={A 2 x 4 factorial arrangement of treatments was used in a randomized complete block designed study to determine the effects of chromium level and source on growth and immune response of stressed and non-stressed 3-wk-old crossbred weanling pigs (BW was 6.35 kg). Factors included 1) immune stress or control and 2) no supplemental Cr or .2 ppm of supplemental Cr from either CrCl3, Cr-picolinate, or Cr-nicotinic acid complex. The basal diet was a corn-soybean meal-whey diet containing 1.2% lysine. Escherichia coli lipopolysaccharide (LPS) was the stress-inducing agent and was injected on d 7, 10, and 13 of the experiment. Immune challenge with LPS resulted in reduced gain (P < .05) and feed intake (P < .10). Supplementation with Cr was not effective in alleviating the depression in growth due to LPS. However, supplementation of control pigs with Cr tended to improve (P < .10) gain and feed intake. In vitro cellular immune response as measured by a lymphocyte blastogenesis assay was increased (P < .10) in pigs fed supplemental Cr from CrCl3, or Cr-picolinate. Antibody response to sheep red blood cells tended to be increased (P < .10) in pigs supplemented with Cr-nicotinic acid, but antibody response to ovalbumin was decreased (P < .05) in pigs supplemented with organic forms of Cr. At the end of the study, effects of Cr supplementation on lymphocyte proliferative response were investigated before and after ACTH administration. Injections of ACTH resulted in increased (P < .001) serum cortisol levels and increased lymphocyte proliferation. Supplementation of Cr did not affect lymphocyte blastogenic response before or after ACTH injection (P > .10). These data suggest that Cr supplementation was not beneficial during immune stress in pigs.}, number={2}, journal={Journal of Animal Science}, author={Heugten, Eric and Spears, J.W.}, year={1997}, pages={409–416} } @article{heugten_coffey_spears_1996, title={Effects of immune challenge, dietary energy density, and source of energy on performance and immunity in weanling pigs}, volume={74}, DOI={10.2527/1996.74102431x}, abstractNote={The objective of this study was to investigate the effects of nutrient density and dietary energy source on performance and immune function of weanling pigs that were either challenged or not challenged with Escherichia coli lipopolysaccharide (LPS). A basal diet was formulated to contain 14 g CP/MJ DE and 7 g lysine/100 g CP. Sulfur amino acids, threonine and tryptophan were kept constant relative to lysine. Experimental diets were mixed using 70 parts basal diet and either 30 parts starch or an isocaloric amount (14 parts) of lard. Diets were fed either for ad libitum intake or on a pair-feeding basis to evaluate effects of diet nutrient density or source of energy, respectively. On d 9 and 25, pigs were challenged i.m. with either 1 mL of a LPS solution or a control solution. Lymphocyte blastogenesis was measured 2 d after the LPS administration and antibody response to sheep red blood cells (SRBC) or ovalbumin was determined 3 d after challenge. No interactive effects on performance were observed between LPS challenge and energy density or source of energy (P > .10). Injection of LPS tended to reduce feed intake and daily gain (P < .10), but not efficiency of feed or energy utilization. Addition of fat to the diets improved feed efficiency and efficiency of energy utilization for gain (P < .05). No consistent effects of LPS challenge, energy density, or source of energy were observed for lymphocyte blastogenesis. Antibody response to ovalbumin, but not to SRBC, was decreased by fat (P < .05). Results indicate that increasing energy density of the diet did not alter the performance depression due to LPS challenge. Addition of fat to the diet improved feed efficiency and efficiency of energy conversion but may depress the humoral immune response. Effects of fat on the immune response may depend on the immune status of the pig.}, number={10}, journal={Journal of Animal Science}, author={Heugten, Eric and Coffey, M. T. and Spears, J. W.}, year={1996}, pages={2431} } @article{vanheugten_spears_coffey_1994, title={THE EFFECT OF DIETARY-PROTEIN ON PERFORMANCE AND IMMUNE-RESPONSE IN WEANLING PIGS SUBJECTED TO AN INFLAMMATORY CHALLENGE}, volume={72}, ISSN={["1525-3163"]}, DOI={10.2527/1994.72102661x}, abstractNote={A total of 96, 21-d-old, crossbred weanling pigs (average initial weight was 6.0 kg) were assigned to one of six treatments to investigate the effect of dietary protein on performance and immune function of Escherichia coli lipopolysaccharide (LPS)-challenged and unchallenged pigs. A control diet was formulated to contain 14.7 MJ of DE/kg, 14 g of CP/MJ of DE, and 7 g of lysine/100 g of CP. Diets low and high in protein were formulated by changing protein levels to 60 or 120% of the control diet. On d 7 and 21, pigs were challenged with either a LPS solution or a saline solution. Lymphocyte blastogenesis was measured 2 d after LPS challenges and antibody response to sheep red blood cells (SRBC) or ovalbumin was measured 3 d after the challenges. Gain and feed consumption were determined 3 d after each LPS injection and at weekly intervals for a total period of 5 wk. Injection of LPS decreased daily gain, feed intake, feed efficiency, and efficiency of protein utilization (P < .05). No interactive effects between LPS challenge and dietary protein were observed for pig performance (P > .10). Daily gain and feed efficiency were improved when protein level was increased from 60 to 100% of the control diet (P < .01). Efficiency of protein utilization for weight gain was lower when the 120% protein diet was fed (P < .01). Antibody response to SRBC or ovalbumin was not affected by treatments.(ABSTRACT TRUNCATED AT 250 WORDS)}, number={10}, journal={JOURNAL OF ANIMAL SCIENCE}, author={VANHEUGTEN, E and SPEARS, JW and COFFEY, MT}, year={1994}, month={Oct}, pages={2661–2669} } @article{vanheugten_spears_coffey_kegley_qureshi_1994, title={THE EFFECT OF METHIONINE AND AFLATOXIN ON IMMUNE FUNCTION IN WEANLING PIGS}, volume={72}, ISSN={["0021-8812"]}, DOI={10.2527/1994.723658x}, abstractNote={To investigate the effect of aflatoxin (AF) and dietary methionine (MET) on immune responses of swine, a total of 288 pigs weaned at 21 d of age were allotted to 12 dietary treatments arranged in a 3 x 4 factorial arrangement in a randomized complete block design. Diets consisted of a corn-soybean meal diet (.95% lysine, .30% MET, and .32% cystine) containing either 0, 140, or 280 ppb of AF and supplemented with either 0, .15, .30, or .45% DL-MET. Immune response measurements were made after the pigs had received their diet for 3 wk. Antibody response to sheep red blood cells (SRBC) was measured 0, 7, and 14 d after i.m. injection of 2.5 mL of a 20% SRBC suspension. Total serum immunoglobulin (Ig) M and IgG were measured using an ELISA. In vivo cellular immunity was measured using a phytohemagglutinin (PHA) skin test. Skin thickness was measured 0, 6, 12, 24, and 36 h after s.c. injection of .1 mL of PHA (1.50 mg/mL). In vitro cellular immunity was measured using a lymphocyte blastogenesis assay. Antibody response to SRBC and serum IgM and IgG concentrations were not affected by dietary treatments. Skin thickness response at 6 h after injection was maximal when .45% MET was added to diets containing 280 ppb of AF, whereas the response was maximal at .30% supplemental MET for the 0 and 140 ppb of AF diets (AF x MET interaction, P < .10). Skin thickness was reduced linearly (P < .10) with increasing dietary AF at 12 and 24 h after PHA injection.(ABSTRACT TRUNCATED AT 250 WORDS)}, number={3}, journal={JOURNAL OF ANIMAL SCIENCE}, author={VANHEUGTEN, E and SPEARS, JW and COFFEY, MT and KEGLEY, EB and QURESHI, MA}, year={1994}, month={Mar}, pages={658–664} } @inproceedings{seventh crissey zoological nutrition symposium, publisher={Raleigh, NC: North Carolina State University, College of Veterinary Medicine}, pages={1–105} }