@article{jennings_wise_nickeleit_maes_cianciolo_piero_law_kim_mccalla_breuhaus_et al._2013, title={Polyomavirus-Associated Nephritis in 2 Horses}, volume={50}, ISSN={0300-9858 1544-2217}, url={http://dx.doi.org/10.1177/0300985813476063}, DOI={10.1177/0300985813476063}, abstractNote={ Polyomaviruses produce latent and asymptomatic infections in many species, but productive and lytic infections are rare. In immunocompromised humans, polyomaviruses can cause tubulointerstitial nephritis, demyelination, or meningoencephalitis in the central nervous system and interstitial pneumonia. This report describes 2 Standardbred horses with tubular necrosis and tubulointerstitial nephritis associated with productive equine polyomavirus infection that resembles BK polyomavirus nephropathy in immunocompromised humans. }, number={5}, journal={Veterinary Pathology}, publisher={SAGE Publications}, author={Jennings, S. H. and Wise, A. G. and Nickeleit, V. and Maes, R. K. and Cianciolo, R. E. and Piero, F. Del and Law, J. M. and Kim, Y. and McCalla, A. C. and Breuhaus, B. A. and et al.}, year={2013}, month={Feb}, pages={769–774} } @article{roberts_2008, title={Equine lymphoma: What are the prospects for cellular differentiation, early diagnosis and intervention strategies?}, volume={20}, ISSN={["2042-3292"]}, DOI={10.2746/095777308x336327}, abstractNote={Equine Veterinary EducationVolume 20, Issue 9 p. 464-466 Equine lymphoma: What are the prospects for cellular differentiation, early diagnosis and intervention strategies? M. C. Roberts, M. C. Roberts College of Veterinary Medicine, North Carolina State University, Raleigh, North Carolina 27606, USA.Search for more papers by this author M. C. Roberts, M. C. Roberts College of Veterinary Medicine, North Carolina State University, Raleigh, North Carolina 27606, USA.Search for more papers by this author First published: 05 January 2010 https://doi.org/10.2746/095777308X336327Citations: 7AboutPDF ToolsExport citationAdd to favoritesTrack citation ShareShare Give accessShare full text accessShare full-text accessPlease review our Terms and Conditions of Use and check box below to share full-text version of article.I have read and accept the Wiley Online Library Terms and Conditions of UseShareable LinkUse the link below to share a full-text version of this article with your friends and colleagues. Learn more.Copy URL No abstract is available for this article.Citing Literature Volume20, Issue9September 2008Pages 464-466 RelatedInformation}, number={9}, journal={EQUINE VETERINARY EDUCATION}, author={Roberts, M. C.}, year={2008}, month={Sep}, pages={464–466} } @article{olivry_jackson_murphy_tater_roberts_2005, title={Evaluation of a point-of-care immunodot assay for predicting results of allergen-specific intradermal and immunoglobulin E serological tests}, volume={16}, ISSN={["0959-4493"]}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-18544380343&partnerID=MN8TOARS}, DOI={10.1111/j.1365-3164.2005.00442.x}, abstractNote={Abstract  Immunotherapy to prevent recurrence of clinical signs of atopic dermatitis (AD) is based on intradermal or serological tests that assist in identifying allergen‐specific immunoglobulin E hypersensitivities. Unfortunately, the results of such tests can be negatively influenced by several factors, which include the age of the patients, the season of testing and the administration of anti‐allergic drugs. Screening to predict when these expensive tests will be useful would benefit owners of dogs with AD. The objectives of this study were to determine whether a point‐of‐care allergen‐specific immunodot assay (Allercept E‐Screen©, Heska Corp., Ft Collins, CO, USA) could predict results of either intradermal or Allercept© full panel serological tests in atopic dogs. Thirty dogs living in the south‐eastern USA were diagnosed with AD in accordance with current standards. Allergen‐specific intradermal, serological and E‐Screen© tests were performed in all subjects. For flea, house dust mite and pollen allergens altogether, results of the E‐Screen© assay agreed with those of intradermal and serological tests in 26/30 dogs (87%) and 25/30 dogs (83%), respectively. In this group of dogs, the probabilities of obtaining intradermal or serological tests positive for these allergens were 70 and 67%, respectively. If either skin or serum tests were performed only in dogs with positive E‐Screen© tests, the probability of obtaining positive results would be increased from 70 to 95% and from 67 to 90%, respectively. In this population of dogs with AD, results of the E‐Screen© point‐of‐care immunodot assay was found to often agree with those of allergen‐specific intradermal or Allercept© tests for selected allergen groups.}, number={2}, journal={VETERINARY DERMATOLOGY}, author={Olivry, T and Jackson, HA and Murphy, KM and Tater, KC and Roberts, M}, year={2005}, month={Apr}, pages={117–120} } @article{sellon_roberts_blikslager_ulibarri_papich_2004, title={Effects of continuous rate intravenous infusion of butorphanol on physiologic and outcome variables in horses after celiotomy}, volume={18}, ISSN={["0891-6640"]}, DOI={10.1892/0891-6640(2004)18<555:EOCRII>2.0.CO;2}, abstractNote={A randomized, controlled, blinded clinical trial was performed to determine whether butorphanol administered by continuous rate infusion (CRI) for 24 hours after abdominal surgery would decrease pain and surgical stress responses and improve recovery in horses. Thirty-one horses undergoing exploratory celiotomy for abdominal pain were randomly assigned to receive butorphanol CRI (13 microg/kg/h for 24 hours after surgery; treatment) or isotonic saline (control). All horses received flunixin meglumine (1.1 mg/kg IV q12h). There were no significant differences between treatment and control horses in preoperative or operative variables. Treatment horses had significantly lower plasma cortisol concentration compared with control horses at 2, 8, 12, 24, 36, and 48 hours after surgery. Mean weight loss while hospitalized was significantly less for treatment horses than control horses, whether expressed as total decrease in body weight (13.9+/-3.4 and 27.9+/-4.5 kg, respectively) or as a percentage decrease in body weight (2.6+/-0.7 and 6.3+/-1.1%, respectively). Treatment horses were significantly delayed in time to first passage of feces (median times of 15 and 4 hours, respectively). Treatment horses had significantly improved behavior scores during the first 24 hours after surgery, consistent with the conclusion that they experienced less pain during that time. Butorphanol CRI during the immediate postoperative period significantly decreased plasma cortisol concentrations and improved recovery characteristics in horses undergoing abdominal surgery.}, number={4}, journal={JOURNAL OF VETERINARY INTERNAL MEDICINE}, author={Sellon, DC and Roberts, MC and Blikslager, AT and Ulibarri, C and Papich, MG}, year={2004}, pages={555–563} } @article{roberts_millikan_galanko_martin_sandler_2003, title={Constipation, laxative use, and colon cancer in a North Carolina population}, volume={98}, ISSN={["0002-9270"]}, DOI={10.1016/S0002-9270(03)00050-9}, number={4}, journal={AMERICAN JOURNAL OF GASTROENTEROLOGY}, author={Roberts, MC and Millikan, RC and Galanko, JA and Martin, C and Sandler, RS}, year={2003}, month={Apr}, pages={857–864} } @article{pritchett_ulibarri_roberts_schneider_sellon_2003, title={Identification of potential physiological and behavioral indicators of postoperative pain in horses after exploratory celiotomy for colic}, volume={80}, ISSN={["0168-1591"]}, DOI={10.1016/S0168-1591(02)00205-8}, abstractNote={Physiological and behavioral parameters were determined in 27 horses to identify potential indicators of postoperative pain following exploratory celiotomy for colic. Experimental groups were 10 horses that received no treatment (Control), 10 horses anesthetized for a non-painful procedure (Anesthesia), and 7 horses that presented for emergency surgery for acute gastrointestinal disease (Surgery). Physiological and behavioral data were collected on the horses 0, 4, 8, 12, 16, 20, and 24–30 h after entry into a stall in the equine intensive care unit of the Veterinary Teaching Hospital at Washington State University. Physiological data included: heart rate, respiratory rate, and plasma cortisol concentration. For the entire period of observation the surgery group had significantly higher plasma cortisol concentration and significantly elevated heart rate compared to the Control and Anesthesia groups, which did not differ for either variable. A numerical rating scale (NRS) of behavior was used to visually score the horses at the same time physiological data were collected. In addition, time budgets of behavior were calculated from 1 h segments of real-time video recording beginning at the 0, 4, 8 or 12 h, and 20 or 24–30 h time points. Time budgets for the Control and Anesthesia groups did not differ in the time spent in locomotor activities and both groups spent significantly more time in locomotion than the Surgery group. The Surgery group spent significantly more time displaying painful behavior compared to the Control and Anesthesia groups; however, the amount of time the Surgery group displayed painful behavior was small compared to the amount of time with no movement. The NRS scores substantiated the video taped behavioral data with significantly different scores for the Surgery group versus the Control and Anesthesia groups for multi-factor ratings of body posture and response to stimuli. We conclude that reduced locomotion, elevated plasma cortisol concentration, and elevated heart rate are potential indicators of postoperative pain in horses.}, number={1}, journal={APPLIED ANIMAL BEHAVIOUR SCIENCE}, author={Pritchett, LC and Ulibarri, C and Roberts, MC and Schneider, RK and Sellon, DC}, year={2003}, month={Jan}, pages={31–43} } @article{mcconnico_argenzio_roberts_2002, title={Prostaglandin E-2 and reactive oxygen metabolite damage in the cecum in a pony model of acute colitis}, volume={66}, number={1}, journal={Canadian Journal of Veterinary Research}, author={McConnico, R. S. and Argenzio, R. A. and Roberts, M. C.}, year={2002}, pages={50–54} } @article{alford_geller_richardson_slater_honnas_foreman_robinson_messer_roberts_goble_et al._2001, title={A multicenter, matched case-control study of risk factors for equine laminitis}, volume={49}, ISSN={["0167-5877"]}, DOI={10.1016/S0167-5877(01)00188-X}, abstractNote={Risk factors for equine laminitis were examined in a prospective case-control study of the 258 cases seen at six collaborating veterinary teaching hospitals over a 32-month period. Case-control pairs were matched on institution, clinician, and season of diagnosis. The 90% of case-control pairs (78 acute, 155 chronic) that had complete data for age, gender, and breed were used in separate conditional logistic-regression models for acute and chronic laminitis. There was an increase in risk for horses with acute laminitis from 5 to 7 years of age (OR 4.7, 95% CI 1.3–16) and from 13 to 31 years of age (OR 3.9, 95% CI 1.3–12) (both compared to <5 years); risk was increased for chronic laminitis from 10 to 14 years (OR 3, 95% CI 1.4–6.8) and from 15 to 38 years (OR 2.9, 95% CI 1.4–6.1) (both compared to <6 years). Mares — but not stallions — were more likely than geldings to develop acute laminitis (OR 2.6, 95% CI 1.1–6.2) and chronic laminitis (OR 2.0, 95% CI 1.1–3.6). In the small acute-laminitis data set, the breed variable was collapsed into three categories: Thoroughbred (THB, reference), the Quarter Horse (QH), and other (non-QH-THB). The non-QH-THB group was at increased risk of acute laminitis (OR 3.8, 95% CI 1.2–11.8). For the seven breed-group categories used in the chronic-laminitis model, however, all non-THB breed groups appeared significantly at risk as compared to the THB, with odds ratios ranging from 3.3 (95% CI 1.3–8.30) for the QH to 9.1 (95% CI 2.1–39.3) for ponies.}, number={3-4}, journal={PREVENTIVE VETERINARY MEDICINE}, author={Alford, P and Geller, S and Richardson, B and Slater, M and Honnas, C and Foreman, J and Robinson, J and Messer, M and Roberts, M and Goble, D and et al.}, year={2001}, month={May}, pages={209–222} } @article{sellon_monroe_roberts_papich_2001, title={Pharmacokinetics and adverse effects of butorphanol administered by single intravenous injection or continuous intravenous infusion in horses}, volume={62}, ISSN={["1943-5681"]}, DOI={10.2460/ajvr.2001.62.183}, abstractNote={Abstract}, number={2}, journal={AMERICAN JOURNAL OF VETERINARY RESEARCH}, author={Sellon, DC and Monroe, VL and Roberts, MC and Papich, MG}, year={2001}, month={Feb}, pages={183–189} } @article{blikslager_roberts_young_rhoads_argenzio_2000, title={Genistein augments prostaglandin-induced recovery of barrier function in ischemia-injured porcine ileum}, volume={278}, ISSN={0193-1857 1522-1547}, url={http://dx.doi.org/10.1152/ajpgi.2000.278.2.G207}, DOI={10.1152/ajpgi.2000.278.2.g207}, abstractNote={We have previously shown that PGE2enhances recovery of transmucosal resistance ( R) in ischemia-injured porcine ileum via a mechanism involving chloride secretion. Because the tyrosine kinase inhibitor genistein amplifies cAMP-induced Cl−secretion, we postulated that genistein would augment PGE2-induced recovery of R. Porcine ileum subjected to 45 min of ischemia was mounted in Ussing chambers, and R and mucosal-to-serosal fluxes of [3H] N-formyl-methionyl-leucyl phenylalanine (FMLP) and [3H]mannitol were monitored as indicators of recovery of barrier function. Treatment with genistein (10−4M) and PGE2(10−6M) resulted in synergistic elevations in R and additive reductions in mucosal-to-serosal fluxes of [3H]FMLP and [3H]mannitol, whereas treatment with genistein alone had no effect. Treatment of injured tissues with genistein and either 8-bromo-cAMP (10−4M) or cGMP (10−4M) resulted in synergistic increases in R. However, treatment of tissues with genistein and the protein kinase C (PKC) agonist phorbol myristate acetate (10−5–10−6M) had no effect on R. Genistein augments recovery of Rin the presence of cAMP or cGMP but not in the presence of PKC agonists.}, number={2}, journal={American Journal of Physiology-Gastrointestinal and Liver Physiology}, publisher={American Physiological Society}, author={Blikslager, Anthony T. and Roberts, Malcolm C. and Young, Karen M. and Rhoads, J. Marc and Argenzio, Robert A.}, year={2000}, month={Feb}, pages={G207–G216} } @article{roberts_2000, title={Small intestinal malabsorption in horses}, volume={12}, DOI={10.1111/j.2042-3292.2000.tb00043.x}, abstractNote={Equine Veterinary EducationVolume 12, Issue 4 p. 214-219 Small intestinal malabsorption in horses M. C. Roberts, M. C. Roberts North Carolina State University, College of Veterinary Medicine, Department of Farm Animal Health and Resource Management, 4700 Hillsborough Street at William Moore Drive, Raleigh, North Carolina 27606–1499, USA.Search for more papers by this author M. C. Roberts, M. C. Roberts North Carolina State University, College of Veterinary Medicine, Department of Farm Animal Health and Resource Management, 4700 Hillsborough Street at William Moore Drive, Raleigh, North Carolina 27606–1499, USA.Search for more papers by this author First published: 05 January 2010 https://doi.org/10.1111/j.2042-3292.2000.tb00043.xCitations: 8AboutPDF ToolsExport citationAdd to favoritesTrack citation ShareShare Give accessShare full text accessShare full-text accessPlease review our Terms and Conditions of Use and check box below to share full-text version of article.I have read and accept the Wiley Online Library Terms and Conditions of UseShareable LinkUse the link below to share a full-text version of this article with your friends and colleagues. Learn more.Copy URL Share a linkShare onEmailFacebookTwitterLinkedInRedditWechat References Brown, C.M. (1992) The diagnostic value of the D-xylose absorption test in horses with unexplained chronic weight loss. Br. vet. J. 148, 41–44. 10.1016/0007-1935(92)90065-9 CASPubMedWeb of Science®Google Scholar Church, S. and Middleton, D.J. (1997) Transient glucose malabsorption in two horses - fact or artifact? Aust. vet. J. 75, 716–718. 10.1111/j.1751-0813.1997.tb12251.x CASPubMedWeb of Science®Google Scholar Freeman, D.E., Ferrante, P.L., Kronfield, D.S. and Chalupa, W. (1989) Effect of food deprivation on D-xylose absorption test results in mares. Am. J. vet. Res. 50, 1609–1612. CASPubMedWeb of Science®Google Scholar Gibson, K.T. and Alders, R.G. (1987) Eosinophilic enterocolitis and dermatitis in two horses. Equine vet. J. 19, 247–252. 10.1111/j.2042-3306.1987.tb01397.x CASPubMedWeb of Science®Google Scholar Haven, M.L. (1994) Effects of Extensive Small Intestinal Resection in the Pony. PhD thesis, North Carolina State University. Google Scholar Lindberg, R., Persson, S.G.B., Jones, B., Thoren-Tolling, K. and Ederoth, M. (1985) Clinical and pathophysiological features of granulomatous enteritis and eosinophilic granulomatosis in the horse. Zbl. vet. Med. A. 32, 526–539. 10.1111/j.1439-0442.1985.tb01973.x CASPubMedWeb of Science®Google Scholar Lindberg, R., Nygren, A. and Persson, S.G.B. (1996) Rectal biopsy diagnosis in horses with clinical signs of intestinal disorders: a retrospective study of 116 cases. Equine vet. J. 28, 275–284. 10.1111/j.2042-3306.1996.tb03091.x CASPubMedWeb of Science®Google Scholar Love, S., Mair, T.S. and Hillyer, M.H. (1992) Chronic diarrhea in adult horses: a review of 51 referred cases. Vet. Rec. 130, 217–219. 10.1136/vr.130.11.217 CASPubMedWeb of Science®Google Scholar Mair, T.S., Hillyer, M.H., Taylor, F.G.R. and Pearson, G.R. (1991) Small intestinal malabsorption in the horse: an assessment of the specificity of the oral glucose tolerance test. Equine vet. J. 23, 344–346. 10.1111/j.2042-3306.1991.tb03735.x CASPubMedWeb of Science®Google Scholar Murphy, D., Reid, S.W.J. and Love, S. (1998) Breath hydrogen measurement in ponies: a preliminary study. Res. vet. Sci. 65, 47–51. 10.1016/S0034-5288(98)90026-1 CASPubMedWeb of Science®Google Scholar Nimmo Wilkie, J.S., Yager, J.A., Nation, P.N., Clark, E.G., Townsend, H.G.G. and Baird, J.D. (1985) Chronic eosinophilic dermatitis: a manifestation of a multisystemic, eosinophilic, epitheliotropic disease in five horses. Vet. Path. 22, 297–305. 10.1177/030098588502200401 CASPubMedWeb of Science®Google Scholar Pass, D.A. and Bolton, J.R. (1982) Chronic eosinophilic gastroenteritis in the horse. Vet. Path. 19, 486–496. 10.1177/030098588201900504 CASPubMedWeb of Science®Google Scholar Roberts, M.C. (1985) Malabsorption syndromes in the horse. Comp. cont. Educ. pract. Vet. 7, S637–S646. Web of Science®Google Scholar Tate, L.P., Ralston, S.L., Koch, C.M. and Everitt, J.E. (1983) Effects of extensive resection of the small intestine in the pony. Am. J. vet. Res. 44, 1187–1191. CASPubMedWeb of Science®Google Scholar Citing Literature Volume12, Issue4June 2000Pages 214-219 ReferencesRelatedInformation}, number={4}, journal={Equine Veterinary Education}, author={Roberts, M. C.}, year={2000}, pages={214–219} } @article{blikslager_rhoads_bristol_roberts_argenzio_1999, title={Glutamine and transforming growth factor-α stimulate extracellular regulated kinases and enhance recovery of villous surface area in porcine ischemic-injured intestine}, volume={125}, url={http://www.scopus.com/inward/record.url?eid=2-s2.0-0033056065&partnerID=MN8TOARS}, DOI={10.1016/S0039-6060(99)70264-3}, abstractNote={

Abstract

Background: Epidermal growth factor (EGF) signals enterocyte proliferation via extracellular regulated kinases (ERKs). Because glutamine is required for EGF-stimulated proliferation and stimulates ERKs in intestinal cell culture, we hypothesized that glutamine and the EGF-related peptide transforming growth factor–alpha (TGF-α) would synergistically enhance repair associated with stimulation of ERKs. Methods: Thiry-Vella loops were created in juvenile pigs. One half of the loop was subjected to 2 hours of ischemia, and the other half served as control. Loops were infused daily with Ringer's solution containing 140 mmol/L glucose, 140 mmol/L glutamine, 140 mmol/L glucose plus 60 μg/L TGF-α, or 140 mmol/L glutamine plus 60 μg/L TGF-α. Results: After 2 hours of ischemia, complete villous epithelial sloughing was present. By 18 hours, villous epithelium had fully restituted, but villi remained stunted until 144 hours after injury. Glutamine + TGF-α triggered sustained increases in ERK activity compared with glucose-treated tissues (maximal at 18 hours), whereas glutamine alone or glucose + TGF-α caused only transient elevations in ERK activity. By 72 hours, villous surface area had increased to normal values with glutamine plus TGF-α treatment, whereas villi remained stunted with glucose alone, glutamine alone, or glucose plus TGF-α. Conclusions: Glutamine plus TGF-α treatment restored mucosal architecture within 72 hours of severe ischemic injury associated with sustained elevations in ERK activity. (Surgery 1999;125:186-94.)}, number={2}, journal={Surgery}, author={Blikslager, Anthony and Rhoads, J.M. and Bristol, D.G. and Roberts, M.C. and Argenzio, R.A.}, year={1999}, pages={186–194} } @article{mcconnico_weinstock_poston_roberts_1999, title={Myeloperoxidase activity of the large intestine in an equine model of acute colitis}, volume={60}, number={7}, journal={American Journal of Veterinary Research}, author={McConnico, R. S. and Weinstock, D. and Poston, M. E. and Roberts, M. C.}, year={1999}, pages={807–813} } @article{blikslager_roberts_argenzio_1999, title={Prostaglandin-induced recovery of barrier function in porcine ileum is triggered by chloride secretion}, volume={276}, ISSN={0193-1857 1522-1547}, url={http://dx.doi.org/10.1152/ajpgi.1999.276.1.G28}, DOI={10.1152/ajpgi.1999.276.1.g28}, abstractNote={ We have previously shown that PGI2 and PGE2 have a synergistic role in restoring electrical transepithelial resistance ( R) in ischemia-injured porcine ileum via the second messengers Ca2+ and cAMP. Because Ca2+ and cAMP stimulate Cl− secretion, we assessed the role of PG-induced Cl−secretion in recovery of R. Mucosa from porcine ileum subjected to ischemia for 45 min was mounted in Ussing chambers and bathed in indomethacin and Ringer solution. Addition of PGs stimulated a twofold increase in R, which was preceded by elevations in short-circuit current (increase of 25 μA/cm2). The PG-induced effect on R was partially inhibited with bumetanide, an inhibitor of Cl− secretion. The remaining elevations in R were similar in magnitude to those induced in ischemic tissues by amiloride, an inhibitor of Na+ absorption. Treatment with 10−4 M 8-bromo-cGMP or 300 mosM mucosal urea resulted in elevations in R similar to those attained with PG treatment. PGs signal recovery of Rvia induction of Cl−secretion and inhibition of Na+absorption, possibly by establishing a transmucosal osmotic gradient. }, number={1}, journal={American Journal of Physiology-Gastrointestinal and Liver Physiology}, publisher={American Physiological Society}, author={Blikslager, Anthony T. and Roberts, Malcolm C. and Argenzio, Robert A.}, year={1999}, month={Jan}, pages={G28–G36} } @inbook{berschneider_blikslager_roberts_1999, title={Role of duodenal reflux in nonglandular gastric ulcer disease of the mature horse}, booktitle={Equine gastric ulceration (Equine veterinary journal supplement)}, publisher={Suffolk, UK: British Equine Veterinary Association}, author={Berschneider, H. M. and Blikslager, A. T. and Roberts, M. C.}, editor={T. S. Mair, P. D. Rossdale and Merritt, A. M.Editors}, year={1999}, month={Apr}, pages={24–29} } @article{gerard_blikslager_roberts_tate_argenzio_1999, title={The characteristics of intestinal injury peripheral to strangulating obstruction lesions in the equine small intestine}, volume={31}, ISSN={["0425-1644"]}, DOI={10.1111/j.2042-3306.1999.tb03826.x}, abstractNote={Summary}, number={4}, journal={EQUINE VETERINARY JOURNAL}, author={Gerard, MP and Blikslager, AT and Roberts, MC and Tate, LP and Argenzio, RA}, year={1999}, month={Jul}, pages={331–335} } @article{cook_papich_roberts_bowman_1998, title={Pharmacokinetics of cisapride in horses after intravenous and rectal administration: Erratum}, volume={59}, number={4}, journal={American Journal of Veterinary Research}, author={Cook, G. and Papich, M. G. and Roberts, M. C. and Bowman, K. F.}, year={1998}, pages={396} } @article{blikslager_roberts_gerard_argenzio_1997, title={How important is intestinal reperfusion injury in horses?}, volume={211}, number={11}, journal={Journal of the American Veterinary Medical Association}, author={Blikslager, A. T. and Roberts, M. C. and Gerard, M. P. and Argenzio, R. A.}, year={1997}, pages={1387–1389} } @article{blikslager_roberts_rhoads_argenzio_1997, title={Is reperfusion injury an important cause of mucosal damage after porcine intestinal ischemia?}, volume={121}, ISSN={["0039-6060"]}, DOI={10.1016/S0039-6060(97)90107-0}, abstractNote={Intestinal ischemic injury is exacerbated by reperfusion in rodent and feline models because of xanthine oxidase-initiated reactive oxygen metabolite formation and neutrophil infiltration. Studies were conducted to determine the relevance of reperfusion injury in the juvenile pig, whose low levels of xanthine oxidase are similar to those of the human being.Ischemia was induced by means of complete mesenteric arterial occlusion, volvulus, or hemorrhagic shock. Injury was assessed by means of histologic examination and measurement of lipid peroxidation. In addition, myeloperoxidase, as a marker of neutrophil infiltration, and xanthine oxidase-xanthine dehydrogenase were measured.Significant ischemic injury was evident after 0.5 to 3 hours of complete mesenteric occlusion or 2 hours of shock or volvulus. In none of these models was the ischemic injury worsened by reperfusion. To maximize superoxide production, pigs were ventilated on 100% O2, but only limited reperfusion injury (1.2-fold increase in histologic grade) was noted. Xanthine oxidase-xanthine dehydrogenase levels were negligible (0.4 +/- 0.4 mU/gm).Reperfusion injury may not play an important role in intestinal injury under conditions of complete mesenteric ischemia and low-flow states in the pig. This may result from low xanthine oxidase-xanthine dehydrogenase levels, which are similar to those found in the human being.}, number={5}, journal={SURGERY}, author={Blikslager, AT and Roberts, MC and Rhoads, JM and Argenzio, RA}, year={1997}, month={May}, pages={526–534} } @article{blikslager_roberts_1997, title={Mechanisms of intestinal mucosal repair}, volume={211}, number={11}, journal={Journal of the American Veterinary Medical Association}, author={Blikslager, A. T. and Roberts, M. C.}, year={1997}, pages={1437–1441} } @article{blikslager_roberts_1997, title={Nitric oxide and the intestinal epithelial barrier: does it protect or damage the gut?}, volume={25}, number={4}, journal={Journal of Pediatric Gastroenterology and Nutrition}, author={Blikslager, A. T. and Roberts, M. C.}, year={1997}, month={Oct}, pages={439–440} } @article{cook_papich_roberts_bowman_1997, title={Pharmacokinetics of cisapride in horses after intravenous and rectal administration}, volume={58}, number={12}, journal={American Journal of Veterinary Research}, author={Cook, G. and Papich, M. G. and Roberts, M. C. and Bowman, K. F.}, year={1997}, pages={1427–1430} } @article{blikslager_roberts_rhoads_argenzio_1997, title={Prostaglandins I2 and E2 have a synergistic role in rescuing epithelial barrier function in porcine ileum.}, volume={100}, ISSN={0021-9738}, url={http://dx.doi.org/10.1172/JCI119723}, DOI={10.1172/JCI119723}, abstractNote={Prostaglandins (PG) are cytoprotective for gastrointestinal epithelium, possibly because they enhance mucosal repair. The objective of the present studies was to assess the role of prostaglandins in intestinal repair. Intestinal mucosa from porcine ileum subjected to 1 h of ischemia was mounted in Ussing chambers. Recovery of normal transepithelial electrical resistance occurred within 2 h, and continued to increase for a further 2 h to a value twice that of control. The latter response was blocked by inhibition of prostaglandin synthesis, and restored by addition of both carbacyclin (an analog of PGI2) and PGE2, whereas the addition of each alone had little effect. Histologically, prostaglandins had no effect on epithelial restitution or villous contraction, indicating that elevations in transepithelial resistance were associated with increases in paracellular resistance. Furthermore, prostaglandin-stimulated elevations in resistance were inhibited with cytochalasin D, an agent known to stimulate cytoskeletal contraction. Synergistic elevations in transepithelial resistance, similar to those of carbacyclin and PGE2, were also noted after treatment with cAMP and A23187 (a calcium ionophore). We conclude that PGE2 and PGI2 have a synergistic role in restoration of intestinal barrier function by increasing intracellular cAMP and Ca2+, respectively, which in turn signal cytoskeletal-mediated tight junction closure.}, number={8}, journal={Journal of Clinical Investigation}, publisher={American Society for Clinical Investigation}, author={Blikslager, A T and Roberts, M C and Rhoads, J M and Argenzio, R A}, year={1997}, month={Oct}, pages={1928–1933} } @article{blikslager_bowman_levine_bristol_roberts_1994, title={Evaluation of factors associated with postoperative ileus in horses: 31 cases (1990-1992)}, volume={205}, number={12}, journal={Journal of the American Veterinary Medical Association}, author={Blikslager, A. T. and Bowman, K. and Levine, J. F. and Bristol, D. G. and Roberts, M. C.}, year={1994}, pages={1748–1752} } @article{roberts_clarke_johnson_1989, title={Castor-oil induced diarrhoea in ponies: A model for acute colitis}, DOI={10.1111/j.2042-3306.1989.tb05658.x}, abstractNote={Summary}, journal={Equine Veterinary Journal}, author={Roberts, M. C. and Clarke, L. L. and Johnson, C. M.}, year={1989}, pages={60} }