@article{greenberg_moorman_elliott_martin_hopey_caldwell_2023, title={Breeding bird abundance and species diversity greatest in high-severity wildfire patches in central hardwood forests}, volume={529}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2022.120715}, abstractNote={In 2016, mixed-severity wildfires in the southern Appalachians created a gradient of forest structures not typical following prescribed burns, providing a unique opportunity to study temporally dynamic conditions and breeding bird response. We measured forest structure and breeding bird communities across a fire-severity gradient in 3 burned and 3 unburned watersheds for 5 years (Y1-Y5). We categorized plots as unburned (NB), low- (L), moderate- (M), or high-severity (H) using a composite fire-severity index. Tree mortality increased with fire-severity category (FSC) and over time; by Y5, 7 % of trees in NB, 11 % in L, 38 % in M, and 71 % in H had died. Shrub recovery was rapid and most pronounced in H, exceeding other FSCs (70 % vs 21 %–44 %) by Y5. Total bird abundance, species richness, and diversity increased over time in H (by Y3) and M (by Y4); by Y5, these metrics were highest in H and twice as high in H as in NB. Low-severity wildfires had no detectable effects on birds. Abundance of 7 species was greatest in higher-severity FSCs; 11 species did not differ among FSC, although ovenbirds (Seiurus aurocapilla) indicated a trend of lower abundance in H. No species was limited to NB, L, or M, whereas disturbance-dependent indigo bunting (Passerina cyanea), chestnut-sided warbler (Setophaga pensylvanica), and eastern towhee (Pipilo erythrophthalmus) were primarily associated with H. Increased richness and diversity were associated with heavy tree mortality and subsequent shrub recovery in H, accompanied by an influx of disturbance-dependent species and positive or neutral responses by most other species. Results highlight the interrelated roles of fire severity and time in driving forest structure and breeding bird response. Breeding birds responded to high-severity burns similarly to silvicultural treatments with heavy canopy reduction documented in other studies, offering possible alternatives when managing for breeding bird diversity in hardwood forests.}, journal={FOREST ECOLOGY AND MANAGEMENT}, publisher={Elsevier BV}, author={Greenberg, Cathryn H. and Moorman, Christopher E. and Elliott, Katherine J. and Martin, Katherine and Hopey, Mark and Caldwell, Peter V}, year={2023}, month={Feb} } @article{casola_peterson_pacifici_sills_moorman_2023, title={Conservation motivations and willingness to pay for wildlife management areas among recreational user groups}, volume={132}, ISSN={["1873-5754"]}, DOI={10.1016/j.landusepol.2023.106801}, abstractNote={Conservation agencies routinely evaluate the costs and benefits of land management and land acquisition options for wildlife management areas (WMAs). Non-market values, for example visitors' consumer surplus, are often absent from these comparisons. Better estimates of willingness to pay (WTP) for WMAs will allow managers to quantify consumer surpluses for different user groups, identify opportunities to generate additional conservation funding, and improve communication with users. We used the contingent valuation method to estimate the WTP for conservation of WMAs by different user groups. We used interval censored regression to estimate WTP for each user group and modeled how WTP varied with visitation frequency, demographics, and type of use. Dual users, those who participated in both licensed (hunting, angling, or trapping) and non-licensed (all other) activities, had greater WTP ($200.07, 95% CI [$161.18, $238.95]) than users who exclusively participated in either a single non-licensed ($74.74, 95% CI [$50.45, $99.02]) or a single licensed activity ($68.21, 95% CI [$48.41, $88.00]). Willingness-to-pay increased with the number of visits to WMAs per year, college education, and income. The most popular donation motivations were that respondents cared about WMA conservation (72%), wanted WMAs to be around for future generations (70%) and personally benefited from the conservation of WMAs (64%). Similar to a scope test, this study demonstrated greater WTP by users who participate in more diverse recreation types on WMAs. Additionally, our findings show that WMA users, particularly users who engage in multiple activities including at least one that does not require a license, enjoy large consumer surpluses and thus could be drawn on for additional financial support for WMA conservation.}, journal={LAND USE POLICY}, author={Casola, William R. and Peterson, M. Nils and Pacifici, Krishna and Sills, Erin O. and Moorman, Christopher E.}, year={2023}, month={Sep} } @article{boone_moorman_moscicki_collier_chamberlain_terando_pacifici_2023, title={Robust assessment of associations between weather and eastern wild turkey nest success}, volume={11}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.22524}, abstractNote={Abstract}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Boone, Wesley W. and Moorman, Christopher E. and Moscicki, David J. and Collier, Bret A. and Chamberlain, Michael J. and Terando, Adam J. and Pacifici, Krishna}, year={2023}, month={Nov} } @article{wightman_ulrey_bakner_cantrell_ruth_rushton_cedotal_kilgo_moscicki_pacifici_et al._2023, title={Survival and cause-specific mortality of male wild turkeys across the southeastern United States}, volume={12}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.22531}, abstractNote={Abstract}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Wightman, Patrick H. and Ulrey, Erin E. and Bakner, Nicholas W. and Cantrell, Jay R. and Ruth, Charles R. and Rushton, Emily and Cedotal, Cody A. and Kilgo, John C. and Moscicki, David J. and Pacifici, Krishna and et al.}, year={2023}, month={Dec} } @article{hausle_forrester_moorman_martin_2023, title={Tradeoffs between timber and wildlife habitat quality increase with density in longleaf pine (Pinus palustris) plantations}, volume={550}, ISSN={["1872-7042"]}, url={https://doi.org/10.1016/j.foreco.2023.121497}, DOI={10.1016/j.foreco.2023.121497}, abstractNote={Longleaf pine (Pinus palustris), a historically abundant tree species in the southeastern United States, is often planted to restore the ecologically and culturally important longleaf pine ecosystem that once covered vast acreages in the southeastern United States. Government cost-share programs that support establishment of these plantations place restrictions on planting rates to promote wildlife habitat, as greater tree planting density may reduce canopy openness and herbaceous plant cover that are critical components of habitat for priority species, including gopher tortoise (Gopherus polyphemus) and Bachman’s sparrow (Peucaea aestivalis). However, there is expressed concern among some forest managers that more open grown trees in the plantations will be of inferior timber quality with more and larger horizontal branches and associated knots. We examined how density affects dynamics and tradeoffs among understory vegetation structure and composition, longleaf pine stem form (branch density and straightness), and longleaf pine survival by sampling 73 plantations of various ages (5–25 years) and planting rates (653–2445 trees per hectare (TPH)/264–990 trees per acre (TPA)) throughout the southeastern United States. We documented a relationship between planting rate and longleaf pine stand density at time of sampling (r = 0.69, p = 0.0001) and relationships between stand density and habitat and timber quality metrics. Greater stand density resulted in lower tree diameters but greater stand basal area than lower stand density. Higher planting rates led to lower branch density and lower straightness grades than lower planting rates. Canopy openness decreased with greater stand density, and bare ground cover and herbaceous cover decreased as density and stand age at time of sampling increased. Based on our results, we suggest that lower maximum planting rates are appropriate when wildlife habitat is a program objective because lower rates result in fewer tradeoffs, as reducing planting rates slows degradation of wildlife habitat when compared with higher maximum planting rates that have only mixed benefits on timber quality.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Hausle, Jacks M. T. and Forrester, Jodi A. and Moorman, Christopher E. and Martin, Melissa R.}, year={2023}, month={Dec} } @article{pharr_cooper_evans_moorman_voss_vukomanovic_marra_2023, title={Using citizen science data to investigate annual survival rates of resident birds in relation to noise and light pollution}, volume={7}, ISSN={["1573-1642"]}, url={https://doi.org/10.1007/s11252-023-01403-2}, DOI={10.1007/s11252-023-01403-2}, abstractNote={Abstract}, journal={URBAN ECOSYSTEMS}, author={Pharr, Lauren D. and Cooper, Caren B. and Evans, Brian and Moorman, Christopher E. and Voss, Margaret A. and Vukomanovic, Jelena and Marra, Peter P.}, year={2023}, month={Jul} } @article{boone_pacifici_moorman_kays_2023, title={Using decoys and camera traps to estimate depredation rates and neonate survival}, volume={18}, ISSN={["1932-6203"]}, DOI={10.1371/journal.pone.0293328}, abstractNote={Ungulate neonates—individuals less than four weeks old—typically experience the greatest predation rates, and variation in their survival can influence ungulate population dynamics. Typical methods to measure neonate survival involve capture and radio-tracking of adults and neonates to discover mortality events. This type of fieldwork is invasive and expensive, can bias results if it leads to neonate abandonment, and may still have high uncertainty about the predator species involved. Here we explore the potential for a non-invasive approach to estimate an index for neonate survival using camera traps paired with decoys that mimic white-tailed deer (Odocoileus virginianus) neonates in the first month of life. We monitored sites with camera traps for two weeks before and after the placement of the neonate decoy and urine scent lure. Predator response to the decoy was classified into three categories: did not approach, approached within 2.5 m but did not touch the decoy, or physically touched the decoy; when conducting survival analyses, we considered these second two categories as dead neonates. The majority (76.3%) of the predators approached the decoy, with 51.1% initiating physical contact. Decoy probability of survival was 0.31 (95% CI = 0.22, 0.35) for a 30-day period. Decoys within the geographic range of American black bear (Ursus americanus) were primarily (75%) attacked by bears. Overall, neonate survival probability decreased as predator abundance increased. The camera-decoy protocol required about ½ the effort and 1/3 the budget of traditional capture-track approaches. We conclude that the camera-decoy approach is a cost-effective method to estimate a neonate survival probability index based on depredation probability and identify which predators are most important.}, number={10}, journal={PLOS ONE}, author={Boone, Hailey M. and Pacifici, Krishna and Moorman, Christopher E. and Kays, Roland}, year={2023}, month={Oct} } @article{moorman_kreh_2022, title={A word from the North Carolina Organizers}, ISSN={["2328-5540"]}, DOI={10.1002/wsb.1308}, abstractNote={Although wild turkey conservation efforts largely are in a post-restoration phase, there is still much to be learned and shared, both among turkey biologists and researchers and with the greater wildlife conservation community. Beginning in 1959, wild turkey researchers and managers have come together every few years from across the United States to share their experiences, present monitoring data and research results, and address novel or timely turkey conservation issues. The other invited paper summarized the potential effects of the Covid-19 pandemic on turkey hunting during spring 2020;the authors surveyed the primary wildlife biologist tasked with wild turkey management for each state to document wild turkey hunter and hunting dynamics before and during the spring 2020 turkey season. [Extracted from the article] Copyright of Wildlife Society Bulletin is the property of Wiley-Blackwell and its content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's express written permission. However, users may print, download, or email articles for individual use. This may be abridged. No warranty is given about the accuracy of the copy. Users should refer to the original published version of the material for the full . (Copyright applies to all s.)}, journal={WILDLIFE SOCIETY BULLETIN}, author={Moorman, Christopher E. and Kreh, Christopher}, year={2022}, month={May} } @article{casola_peterson_sills_pacifici_moorman_2022, title={Economic contributions of wildlife management areas in North Carolina}, volume={140}, ISSN={["1872-7050"]}, DOI={10.1016/j.forpol.2022.102747}, abstractNote={Wildlife management areas (WMAs) provide a wide range of ecosystem services. Among these services, hunting and fishing often make the most obvious contribution to local and state economies through the expenditures of the hunters and anglers. However, the total economic contributions of WMAs also include other forms of recreation that are generally less visible, unlicensed, and less well understood. Quantifying the size of the economic contribution from all recreationists can inform decisions about investment in and management of public lands. To this end, we estimated the direct, indirect, and induced economic contributions of recreation on protected land managed by the state of North Carolina (NC) primarily for hunting, fishing, and wildlife conservation (hereafter WMAs). We collected data on visitation and conducted in-person intercept surveys at 9 WMAs to estimate the number of visits and expenditures per visit for people engaged in activities that required licenses (e.g., hunting) and activities that did not (e.g., hiking and bird watching). We estimated annual visitation on the 9 study WMAs, accounting for differences in location, hunting season, day of the week, and weather. We then predicted annual visitation at all 94 WMAs in NC using a predictive regression model. Most visitors did not engage in any licensed activities, and those visitors spent more per trip on average ($119.83) and had greater variability in expenses than visitors engaged in licensed activities ($84.19). We used the estimates of total annual visits, expenditures per visit, and the distribution of those expenditures across sectors to calculate the economic contribution of recreation on each of the 9 study WMAs and on the entire WMA system in NC. Recreation was responsible for approximately 2200 jobs, $84 million USD in annual labor income, and $140 million USD in value added annually in NC. The majority of this contribution was due to visits made by users not engage in licensed uses of WMAs, as those users were more numerous, spent more per trip, and were more likely to visit WMAs in peri-urban areas with more economic linkages than rural areas.}, journal={FOREST POLICY AND ECONOMICS}, author={Casola, William R. and Peterson, M. Nils and Sills, Erin O. and Pacifici, Krishna and Moorman, Christopher E.}, year={2022}, month={Jul} } @article{lashley_chitwood_dykes_deperno_moorman_2022, title={Human‐mediated trophic mismatch between fire, plants and herbivores}, volume={2022}, ISSN={0906-7590 1600-0587}, url={http://dx.doi.org/10.1111/ecog.06045}, DOI={10.1111/ecog.06045}, abstractNote={Trophic mismatches are commonly reported across a wide array of taxa and can have important implications for species participating in the interaction. However, to date, examples of trophic mismatch have centrally focused on those induced by shifts in climate. Here we report on the potential for humans to induce trophic mismatch by shifting the phenology of fire. Globally, anthropogenic fire ignitions are phenologically mismatched to that of historic lightning ignitions but the effects of this phenological mismatch on trophic interactions are poorly understood. Using fire records from 1980 to 2016 from the southeastern USA, a hotspot of anthropogenic fire, we demonstrate that there is a temporal mismatch between anthropogenic and lightning lit fires in this region. The peak of anthropogenic ignitions (i.e. 45% during March and April) occurred 3 months earlier than the peak in lightning‐ignited fires (i.e. 44% occurred during June and July), a pattern consistent with reports from several other regions and continents. We demonstrate with a field experiment conducted at a nutrient‐poor site in the southeastern U.S., that anthropogenic fire phenology shifts nutrient pulses in resprouting plants so that they mismatch herbivore reproductive demands. Consequently, plant nutrient quality in four commonly consumed forages was below the threshold to meet lactation requirements. Neonates subsequently were more likely to starve when born far from areas burned during the peak month of lightning fire phenology. Our data indicate that human activities may be an additional causative agent of trophic mismatch.}, number={3}, journal={Ecography}, publisher={Wiley}, author={Lashley, Marcus A. and Chitwood, M. Colter and Dykes, Jacob L. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2022}, month={Feb} } @article{powell_buehler_moorman_zobel_harper_2022, title={Vegetation structure and food availability following disturbance in recently restored early successional plant communities}, ISSN={["2328-5540"]}, DOI={10.1002/wsb.1372}, abstractNote={Abstract}, journal={WILDLIFE SOCIETY BULLETIN}, author={Powell, Bonner L. and Buehler, David A. and Moorman, Christopher E. and Zobel, John M. and Harper, Craig A.}, year={2022}, month={Oct} } @article{choi_fish_moorman_deperno_schillaci_2021, title={Breeding-season Survival, Home-range Size, and Habitat Selection of Female Bachman's Sparrows}, volume={20}, ISSN={["1938-5412"]}, url={https://doi.org/10.1656/058.020.0112}, DOI={10.1656/058.020.0112}, abstractNote={Abstract Peucaea aestivalis (Bachman's Sparrow) is a declining songbird endemic to the southeastern US, but lack of basic life-history information for females, including a description of habitat selection, limits effective management. We investigated survival, home-range size, and habitat selection of female Bachman's Sparrows during the breeding season at Fort Bragg Military Installation, NC. We attached radio-transmitters to female sparrows between April and June in 2014–2016 and recorded locations of females every 2–4 days. We estimated seasonal survival and home-range size and, in 2016, we modeled habitat selection of female sparrows within their home range. Estimated breeding-season (90 days) survival (0.941) was greater than a published estimate from South Carolina (0.794), and home-range size (1.48 ha, SE = 0.16) was similar to a published estimate for females and multiple published estimates for male sparrows (min–max = 1–5 ha). Females selected habitat patches with greater woody vegetation and intermediate grass densities than at random locations, suggesting that woody vegetation provides escape and nesting cover for female sparrows. Survival, home-range size, and habitat selection of female Bachman's Sparrows did not differ substantially from males in other studies. Therefore, management focused on male sparrows may concurrently conserve habitat requirements for females.}, number={1}, journal={SOUTHEASTERN NATURALIST}, publisher={BioOne}, author={Choi, Daniel Y. and Fish, Alexander C. and Moorman, Christopher E. and DePerno, Christopher S. and Schillaci, Jessica M.}, year={2021}, month={Mar}, pages={105–116} } @article{rosche_moorman_kroeger_pacifici_jones_deperno_2021, title={Effects of Prescribed Fire on Northern Bobwhite Nesting Ecology}, volume={45}, ISSN={2328-5540 2328-5540}, url={http://dx.doi.org/10.1002/wsb.1183}, DOI={10.1002/wsb.1183}, abstractNote={ABSTRACT}, number={2}, journal={Wildlife Society Bulletin}, publisher={Wiley}, author={Rosche, Sarah B. and Moorman, Christopher E. and Kroeger, Anthony J. and Pacifici, Krishna and Jones, Jeffrey G. and DePerno, Christopher S.}, year={2021}, month={Jun}, pages={249–257} } @article{arthur_varner_lafon_alexander_dey_harper_horn_hutchinson_keyser_lashley_et al._2021, title={Fire Ecology and Management in Eastern Broadleaf and Appalachian Forests}, volume={39}, ISBN={["978-3-030-73266-0"]}, ISSN={["1568-1319"]}, DOI={10.1007/978-3-030-73267-7_4}, abstractNote={The role of fire in the eastern broadleaf and Appalachian forest regions, until recently, was poorly understood or minimally examined, as this region was long overlooked as a flammable landscape and fire was seen primarily as a threat to the timber resource and wildlife. In the past few decades, a significant body of research has enhanced our understanding of fire and its effects. We now recognize that fire has strongly shaped many ecosystems of this region along complex geomorphological gradients, and that returning fire, or its absence, has significant consequences for forest structure, species composition, and ecosystem function. This chapter synthesizes the state of knowledge regarding the prehistoric and historical roles of fire in these ecosystems; presents research demonstrating the effects of contemporary prescribed fire and wildfire on forest structure, species composition, and consequences for wildlife; examines evidence for shifting flammability of these ecosystems; and discusses the implications for both fire management and ecosystem sustainability in the twenty-first century.}, journal={FIRE ECOLOGY AND MANAGEMENT PAST, PRESENT, AND FUTURE OF US FORESTED ECOSYSTEMS}, author={Arthur, Mary A. and Varner, J. Morgan and Lafon, Charles W. and Alexander, Heather D. and Dey, Daniel C. and Harper, Craig A. and Horn, Sally P. and Hutchinson, Todd F. and Keyser, Tara L. and Lashley, Marcus A. and et al.}, year={2021}, pages={105–147} } @article{harper_gefellers_buehler_moorman_zobel_2021, title={Plant Community Response and Implications for Wildlife Following Control of a Nonnative Perennial Grass}, ISSN={["2328-5540"]}, DOI={10.1002/wsb.1232}, abstractNote={ABSTRACT}, journal={WILDLIFE SOCIETY BULLETIN}, author={Harper, Craig A. and Gefellers, J. Wade and Buehler, David A. and Moorman, Christopher E. and Zobel, John M.}, year={2021}, month={Dec} } @article{casola_peterson_pacifici_moorman_2021, title={Public support and visitation impacts of Sunday hunting on public hunting lands}, volume={26}, ISSN={["1533-158X"]}, DOI={10.1080/10871209.2020.1811923}, abstractNote={Sunday hunting legislation is complex, and often controversial, resulting in recreation impacts for both traditional (hunters, anglers, trappers) and non-traditional (e.g., hikers, birders, bikers,...}, number={1}, journal={HUMAN DIMENSIONS OF WILDLIFE}, author={Casola, William R. and Peterson, M. Nils and Pacifici, Krishna and Moorman, Christopher E.}, year={2021}, month={Jan}, pages={94–97} } @article{anderson_ury_taillie_ungberg_moorman_poulter_ardon_bernhardt_wright_2021, title={Salinity thresholds for understory plants in coastal wetlands}, volume={11}, ISSN={["1573-5052"]}, DOI={10.1007/s11258-021-01209-2}, abstractNote={The effects of sea level rise and coastal saltwater intrusion on wetland plants can extend well above the high-tide line due to drought, hurricanes, and groundwater intrusion. Research has examined how coastal salt marsh plant communities respond to increased flooding and salinity, but more inland coastal systems have received less attention. The aim of this study was to identify whether ground layer plants exhibit threshold responses to salinity exposure. We used two vegetation surveys throughout the Albemarle-Pamlico Peninsula (APP) of North Carolina, USA to assess vegetation in a low elevation landscape (≤ 3.8 m) experiencing high rates of sea level rise (3–4 mm/year). We examined the primary drivers of community composition change using Non-metric Multidimensional Scaling (NMDS) and used Threshold Indicator Taxa Analysis (TITAN) to detect thresholds of compositional change based on indicator taxa, in response to potential indicators of exposure to saltwater (Na, and the Σ Ca + Mg) and elevation. Salinity and elevation explained 64% of the variation in community composition, and we found two salinity thresholds for both soil Na+ (265 and 3843 g Na+/g) and Ca+ + Mg+ (42 and 126 µeq/g) where major changes in community composition occur on the APP. Similar sets of species showed sensitivity to these different metrics of salt exposure. Overall, our results showed that ground layer plants can be used as reliable indicators of salinity thresholds in coastal wetlands. These results can be used for monitoring salt exposure of ecosystems and for identifying areas at risk for undergoing future community shifts.}, journal={PLANT ECOLOGY}, author={Anderson, Steven M. and Ury, Emily A. and Taillie, Paul J. and Ungberg, Eric A. and Moorman, Christopher E. and Poulter, Benjamin and Ardon, Marcelo and Bernhardt, Emily S. and Wright, Justin P.}, year={2021}, month={Nov} } @article{hannon_moorman_schultz_deperno_2021, title={The relationship between upland hardwood distribution and avian occupancy in fire-maintained longleaf pine forests}, volume={479}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/j.foreco.2020.118546}, DOI={10.1016/j.foreco.2020.118546}, abstractNote={Prescribed fire and other forest management practices aimed at restoring longleaf pine (Pinus palustris) communities often focus on the reduction, or removal, of upland hardwoods with the goal of providing habitat for threatened and endangered plant and animal species, including the federally endangered red-cockaded woodpecker (Leuconotopicus borealis), and restoring forests to pre-settlement conditions. Although contemporary restoration and management practices benefit species dependent on the resulting conditions, recent research has called attention to the ecological value of retaining upland hardwoods, especially for mast-dependent wildlife (e.g., fox squirrels [Sciurus niger]). Moreover, retention of indigenous hardwoods in upland longleaf pine communities may benefit a variety of birds. We used fixed-radius, breeding season point counts to sample the presence-absence of 15 avian species and assessed forest composition and structure around each point. We developed single-season single-species occupancy models with an emphasis on the influence of overstory hardwood cover on occupancy. Due to issues with model fit, we were unable to model occupancy for 3 of the 15 focal species. Occupancy probabilities for 6 out of the 12 focal species were positively influenced by overstory hardwood cover or stem density, whereas occupancy probabilities of 4 out of 12 of the focal species was negatively influenced by hardwood cover or stem density. Overstory hardwood cover between 5 and 15% resulted in high occupancy probabilities for the species that were positively influenced but did not result in substantially low occupancy probabilities for the species that were negatively influenced. Longleaf pine uplands with lower and upper bounds of 5% to 15% hardwood overstory cover with hardwood stem densities of ≤250 stems/ha could be targeted to provide habitat for the greatest diversity of birds while avoiding negative impact to species associated with upland longleaf pine communities.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Hannon, Daniel R. and Moorman, Christopher E. and Schultz, Alan D. and DePerno, Christopher S.}, year={2021}, month={Jan}, pages={118546} } @article{noonan_fleming_tucker_kays_harrison_crofoot_abrahms_alberts_ali_altmann_et al._2020, title={Effects of body size on estimation of mammalian area requirements}, volume={34}, ISSN={["1523-1739"]}, url={https://doi.org/10.1111/cobi.13495}, DOI={10.1111/cobi.13495}, abstractNote={Abstract}, number={4}, journal={CONSERVATION BIOLOGY}, author={Noonan, Michael J. and Fleming, Christen H. and Tucker, Marlee A. and Kays, Roland and Harrison, Autumn-Lynn and Crofoot, Margaret C. and Abrahms, Briana and Alberts, Susan C. and Ali, Abdullahi H. and Altmann, Jeanne and et al.}, year={2020}, month={Aug}, pages={1017–1028} } @article{garabedian_moorman_peterson_kilgo_2020, title={Effects of group size and group density on trade-offs in resource selection by a group-territorial central-place foraging woodpecker}, volume={162}, ISSN={["1474-919X"]}, DOI={10.1111/ibi.12733}, abstractNote={Trade‐offs in resource selection by central‐place foragers are driven by the need to balance the benefits of selecting resources against the costs of travel from the central place. For group‐territorial central‐place foraging birds, trade‐offs in resource selection are likely to be complicated by a competitive advantage for larger groups at high group density that may limit accessibility of high‐quality distant resources to small groups. We used the group‐territorial, central‐place foraging Red‐cockaded Woodpecker Leuconotopicus borealis (RCW) as a case study to test predictions that increases in group density lead to differences in foraging distances and resource selection for groups of different sizes. We used GPS tracking and LiDAR‐derived habitat data to model effects of group size on foraging distances and selection for high‐quality pines (≥ 35.6 cm diameter at breast height (dbh)) and lower quality pines (25.4–35.6 cm dbh) by RCW groups across low (n = 14), moderate (n = 10) and high group density (n = 10) conditions. At low and moderate group density, all RCW groups selected distant high‐quality pines in addition to those near the central place because competition for resources was low. In contrast, at high group density, larger groups travelled further to select high‐quality pines, whereas smaller groups selected high‐quality pines only when they were close to the central place and, conversely, were more likely to select lower quality pines at greater distances from the central place. Selection for high‐quality pines only when close to the cavity tree cluster at high group density is important to long‐term fitness of small RCW groups because it allows them to maximize benefits from both territorial defence and selecting high‐quality resources while minimizing costs of competition. These relationships suggest that intraspecific competition at high group density entails substantive costs to smaller groups of territorial central‐place foragers by limiting accessibility of distant high‐quality foraging resources.}, number={2}, journal={IBIS}, author={Garabedian, James E. and Moorman, Christopher E. and Peterson, M. Nils and Kilgo, John C.}, year={2020}, month={Apr}, pages={477–491} } @article{fish_deperno_schillaci_moorman_2020, title={Fledgling Bachman's Sparrows in a longleaf pine ecosystem: survival, movements, and habitat selection}, volume={91}, ISSN={["1557-9263"]}, DOI={10.1111/jofo.12345}, abstractNote={Fledgling ecology remains understudied for many passerine species, yet information about the fledgling life stage is critical for understanding full-annual life cycles and population recruitment. We examined the survival, habitat selection, and movements of fledgling Bachman’s Sparrows (Peucaea aestivalis) in a longleaf pine-wiregrass (Pinus palustris-Aristida stricta) community managed with frequent prescribed fire. We captured and marked 36 fledglings on the day of fledging and used radio-telemetry to relocate them daily until independence during three breeding seasons (2014–2016). We visually confirmed the status of fledglings as live or dead during daily relocations and determined causes of mortality. We measured vegetation characteristics at fledgling locations and compared them to the characteristics of vegetation at the locations of adult males. We used a Known Fates analysis in Program MARK to estimate fledgling survival, and generalized linear mixed effect models to determine habitat selection. Estimated fledgling survival until independence was 0.31 (SE = 0.08), with most mortality during the first 4 d post-fledging. Fledglings with longer wing chords had higher rates of survival than those with shorter wing chords, possibly due to an increased ability to evade predators. Fledgling movements were restricted primarily to natal territories. Fledgling Bachman’s Sparrows were located in areas with greater woody plant, forb, and grass cover and less bare ground than available in natal territories. Similar to fledglings of other songbirds, understory woody and herbaceous plants appear to provide critical cover for fledgling Bachman’s Sparrows, and maintenance of such cover should receive consideration in management plans for longleaf pine communities. RESUMEN. Volantones de Chingolo de Bachman en un ecosistema de pinos de hoja larga: supervivencia, movimientos y selecci on de h abitat La ecolog ıa de los volantones permanece poco estudiada para muchas especies de paseriformes, sin embargo, la informaci on sobre la etapa volantona de la vida es fundamental para comprender los ciclos de vida anuales completos y el reclutamiento de la poblaci on. Examinamos la supervivencia, la selecci on del h abitat y los movimientos del Chingolo de Bachman (Peucaea aestivalis) en una comunidad de pino hoja larga-pasto alambre (Pinus palustris Aristida stricta) manejada con frecuentes incendios controlados. Capturamos y marcamos 36 volantones el d ıa del abandono del nido y utilizamos radiotelemetr ıa para reubicarlos diariamente hasta la independencia durante tres temporadas de cr ıa (2014–2016). Confirmamos visualmente el estado de los volantones como vivos o muertos durante las reubicaciones diarias y determinamos las causas de mortalidad. Medimos las caracter ısticas de la vegetaci on en las ubicaciones de los volantones y las comparamos con las caracter ısticas de la vegetaci on en las ubicaciones de los machos adultos. Usamos un an alisis de Destinos Conocidos en el Programa MARK para estimar la supervivencia de los volantones, y modelos lineales generalizados de efectos mixtos para determinar la selecci on del h abitat. La supervivencia estimada de los volantones hasta la independencia fue de 0.31 (EE = 0.08), con la mayor mortalidad durante los primeros cuatro d ıas despu es del abandono del nido. Los polluelos con cuerdas alares m as larga ten ıan tasas m as altas de supervivencia que aquellos con cuerdas alarer alas m as cortas, posiblemente debido a una mayor capacidad para evadir a los depredadores. Los movimientos de volantones se restringieron principalmente a los territorios natales. Los volantones de Chingolo de Bachman se ubicaron en areas con mayor cobertura de plantas le~ nosas, hierbas y pastos y menos terreno desnudo que el disponible en los territorios natales. Al igual que los polluelos de otras aves cantoras, las plantas le~ nosas y herb aceas del sotobosque parecen proporcionar una cobertura cr ıtica para los volantones de Chingolo de Bachman, y el mantenimiento de dicha cobertura deber ıa ser considerado en los planes de gesti on para las comunidades de pinos de hoja larga.}, number={4}, journal={JOURNAL OF FIELD ORNITHOLOGY}, author={Fish, Alexander C. and DePerno, Christopher S. and Schillaci, Jessica M. and Moorman, Christopher E.}, year={2020}, month={Dec}, pages={354–366} } @article{grodsky_hernandez_campbell_hinson_keller_fritts_homyack_moorman_2020, title={Ground Beetle (Coleoptera: Carabidae) Response to Harvest Residue Retention: Implications for Sustainable Forest Bioenergy Production}, volume={11}, ISSN={["1999-4907"]}, DOI={10.3390/f11010048}, abstractNote={Research Highlights: Our study adds to the scant literature on the effects of forest bioenergy on ground beetles (Coleoptera: Carabidae) and contributes new insights into the responses of ground beetle species and functional groups to operational harvest residue retention. We discovered that count of Harpalus pensylvanicus (DeGeer)—a habitat generalist—increased owing to clear-cut harvests but decreased due to harvest residue reductions; these observations uniquely allowed us to separate effects of additive forest disturbances to demonstrate that, contrarily to predictions, a generalist species considered to be adapted to disturbance may be negatively affected by altered habitat elements associated with disturbances from renewable energy development. Background and Objectives: Despite the potential environmental benefits of forest bioenergy, woody biomass harvests raise forest sustainability concerns for some stakeholders. Ground beetles are well established ecological indicators of forest ecosystem health and their life history characteristics are connected to habitat elements that are altered by forest harvesting. Thus, we evaluated the effects of harvest residue retention following woody biomass harvest for forest bioenergy on ground beetles in an operational field experiment. Materials and Methods: We sampled ground beetles using pitfall traps in harvest residue removal treatments representing variable woody biomass retention prescriptions, ranging from no retention to complete retention of all merchantable woody biomass. We replicated treatments in eight clear-cut stands in intensively managed loblolly pine (Pinus taeda L.) forests in North Carolina and Georgia. Results: Harvest residue retention had no effect on ground beetle richness and diversity. However, counts of H. pensylvanicus, Anisodactylus spp., and “burrower” and “fast runner” functional groups, among others, were greater in treatments with no woody biomass harvest than those with no harvest residue retention; all of these ground beetles may confer ecosystem services in forests. We suggest that H. pensylvanicus is a useful indicator species for burrowing and granivorous ground beetle response to harvest residue reductions in recently harvested stands. Lastly, we propose that retaining 15% retention of total harvest residues or more, depending on regional and operational variables, may support beneficial ground beetle populations.}, number={1}, journal={FORESTS}, author={Grodsky, Steven M. and Hernandez, Rebecca R. and Campbell, Joshua W. and Hinson, Kevin R. and Keller, Oliver and Fritts, Sarah R. and Homyack, Jessica A. and Moorman, Christopher E.}, year={2020}, month={Jan} } @article{boggs_moorman_hazel_greenberg_sorger_sorenson_2020, title={Ground-Dwelling Invertebrate Abundance Positively Related to Volume of Logging Residues in the Southern Appalachians, USA}, volume={11}, ISSN={["1999-4907"]}, DOI={10.3390/f11111149}, abstractNote={Invertebrates, especially those dependent on woody debris for a portion of their life cycle, may be greatly impacted by the amount of downed wood retained following timber harvests. To document relationships between invertebrates and logging residues, we sampled invertebrates with pitfall traps placed near or far from woody debris in 10 recently (2013–2015) harvested sites in western North Carolina with varying levels of woody debris retention. We measured the groundcover and microclimate at each trap and estimated site-level woody debris volume. We modeled predictors (e.g., site-level woody debris volume, percent woody debris cover at the trap site, site type) of captures of spiders (Araneae), harvestmen (Opiliones), centipedes/millipedes (Chilopoda/Diplopoda), ground beetles (Carabidae), rove beetles (Staphylinidae), other beetles, ants (Formicidae), grasshoppers (Acrididae/Tetrigidae), crickets (Gryllidae), and cave crickets (Rhaphidophoridae). In addition, we modeled ant occurrence at a finer taxonomic resolution, including red imported fire ants (Solenopsis invicta Buren) and 13 other genera/species. Forest type, whether hardwood or white pine (Pinus strobus L.) overstory preharvest, was a predictor of invertebrate response for 21 of 24 taxonomic analyses. Invertebrate captures or the occurrence probability of ants increased with increasing site-level woody debris volume for 13 of the 24 taxa examined and increased with increasing coarse woody debris (CWD; diameter ≥ 10 cm) cover at the trap level for seven of 24 taxa examined. Our results indicate that woody debris in harvested sites is important for the conservation of a majority of the taxa we studied, which is likely because of the unique microclimate offered near/under woody debris. Stand-scale factors typically were more important predictors of invertebrate response than trap-level cover of woody debris. We recommend implementing sustainability strategies (e.g., Biomass Harvesting Guidelines) to retain woody debris scattered across harvested sites to aid in the conservation of invertebrates.}, number={11}, journal={FORESTS}, author={Boggs, April D. and Moorman, Christopher E. and Hazel, Dennis W. and Greenberg, Cathryn H. and Sorger, D. Magdalena and Sorenson, Clyde E.}, year={2020}, month={Nov} } @article{drake_peterson_griffith_olfenbuttel_deperno_moorman_2020, title={How Urban Identity, Affect, and Knowledge Predict Perceptions About Coyotes and Their Management}, volume={33}, ISSN={["1753-0377"]}, DOI={10.1080/08927936.2020.1694302}, abstractNote={ABSTRACT Globally, the number of humans and wildlife species sharing urban spaces continues to grow. As these populations grow, so too does the frequency of human–wildlife interactions in urban areas. Carnivores in particular pose urban wildlife conservation challenges owing to the strong emotions they elicit and the potential threats they can present to humans. These challenges can be better addressed with an understanding of the different factors that influence public perceptions of carnivores and their management. We conducted mail surveys in four cities in North Carolina (n =721) to explore how (a) city of residence, (b) affectual connections to coyotes (Canis latrans), and (c) biological knowledge predicted perceptions of the danger posed by coyotes, the support for wild coyotes living nearby, and the support for lethal coyote removal methods. Our results provide the first assessment of how public perceptions of carnivores and their management vary between cities of different types. Residents from a tourism-driven city were more supportive of coyotes than residents from an industrial city and less concerned about risk than residents from a commercial city. We found affectual connection to coyotes and city of residence were consistent predictors of coyote perceptions. Respondents’ knowledge of coyote biology was not a significant predictor of any perceptions of coyotes despite the relatively high statistical power of the tests. Affectual connection to coyotes had the greatest effect on predicting coyote perceptions, suggesting efforts to promote positive emotional connections to wildlife may be a better way to increase acceptance of carnivores in urban areas than focusing on biological knowledge.}, number={1}, journal={ANTHROZOOS}, author={Drake, Michael D. and Peterson, M. Nils and Griffith, Emily H. and Olfenbuttel, Colleen and DePerno, Cristopher S. and Moorman, Christopher E.}, year={2020}, month={Jan}, pages={5–19} } @article{richardson_kroeger_moorman_harper_gardner_jones_strope_2020, title={Nesting Ecology of Northern Bobwhite on a Working Farm}, volume={44}, ISSN={["2328-5540"]}, DOI={10.1002/wsb.1125}, abstractNote={ABSTRACT}, number={4}, journal={WILDLIFE SOCIETY BULLETIN}, author={Richardson, Andy D. and Kroeger, Anthony J. and Moorman, Christopher E. and Harper, Craig A. and Gardner, Beth and Jones, Mark D. and Strope, Benjy M.}, year={2020}, month={Dec}, pages={677–683} } @article{kroeger_deperno_harper_rosche_moorman_2020, title={Northern Bobwhite Non‐Breeding Habitat Selection in a Longleaf Pine Woodland}, volume={84}, ISSN={0022-541X 1937-2817}, url={http://dx.doi.org/10.1002/jwmg.21925}, DOI={10.1002/jwmg.21925}, abstractNote={ABSTRACT}, number={7}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Kroeger, Anthony J. and DePerno, Christopher S. and Harper, Craig A. and Rosche, Sarah B. and Moorman, Christopher E.}, year={2020}, month={Jul}, pages={1348–1360} } @article{hannon_moorman_schultz_gray_deperno_2020, title={Predictors of fire-tolerant oak and fire-sensitive hardwood distribution in a fire-maintained longleaf pine ecosystem}, volume={477}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/j.foreco.2020.118468}, DOI={10.1016/j.foreco.2020.118468}, abstractNote={The longleaf pine (Pinus palustris) ecosystem has been reduced to a fraction of its original extent, and where this ecosystem does occur, it is often degraded by hardwood encroachment. The reduction of hardwood tree cover is often a desirable longleaf pine community restoration outcome, though hardwood midstory and overstory trees have been recognized as an important natural component of the communities. Moreover, the appropriate amount of hardwood tree cover in a restored longleaf pine community is debated, as more hardwood tree cover can benefit mixed forest and mast-dependent wildlife (e.g., fox squirrels [Sciurus niger], white-tailed deer [Odocoileus virginianus]), and less hardwood tree cover is critical to the federally endangered red-cockaded woodpecker (Leuconotopicus borealis). To inform the debate, we assessed the environmental (e.g., topography, edaphic conditions, and pine basal area) and management (e.g., distance to firebreaks, prescribed fire history) factors that influenced abundance of upland hardwood trees in xeric longleaf pine communities on a site where frequent growing-season fire has been ongoing since 1991. We counted upland hardwoods ≥5 cm diameter at breast height (DBH) at 307 random field plots (0.04 ha) and categorized all hardwood trees as belonging to either a guild of fire-tolerant oaks or a guild of fire-sensitive hardwood species. We used generalized linear models (GLM) to determine the most important predictors of abundance for both guilds. The predictors of abundance differed between the two guilds, with fire-tolerant oak abundance increasing with greater slope and proximity to ignition sources and decreasing with greater pine basal area. Fire-sensitive hardwood abundance increased with mesic site conditions and decreased with the number of growing-season fires and greater pine basal area. Although seasonality in fire history was an important predictor of fire-sensitive hardwood abundance, variables related to long-term fire-history were not important predictors of fire-tolerant oak abundance in longleaf pine communities. However, with limited variation in fire return interval across the study area, our ability to draw inferences regarding the role of fire return interval was limited. Where hardwood encroachment is not a problem, and hardwood levels are below desired, balanced target levels, hardwood abundance in longleaf pine communities can be increased by reducing pine basal area and reducing prescribed fire intensity.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Hannon, Daniel R. and Moorman, Christopher E. and Schultz, Alan D. and Gray, Josh M. and DePerno, Christopher S.}, year={2020}, month={Dec}, pages={118468} } @article{chitwood_lashley_higdon_deperno_moorman_2020, title={Raccoon Vigilance and Activity Patterns When Sympatric with Coyotes}, volume={12}, ISSN={1424-2818}, url={http://dx.doi.org/10.3390/d12090341}, DOI={10.3390/d12090341}, abstractNote={Nonconsumptive effects of predators potentially have negative fitness consequences on prey species through changes in prey behavior. Coyotes (Canis latrans) recently expanded into the eastern United States, and raccoons (Procyon lotor) are a common mesocarnivore that potentially serve as competitors and food for coyotes. We used camera traps at baited sites to quantify vigilance behavior of feeding raccoons and used binomial logistic regression to analyze the effects of social and environmental factors. Additionally, we created raccoon and coyote activity patterns from the camera trap data by fitting density functions based on circular statistics and calculating the coefficient of overlap (Δ). Overall, raccoons were vigilant 46% of the time while foraging at baited sites. Raccoons were more vigilant during full moon and diurnal hours but less vigilant as group size increased and when other species were present. Raccoons and coyotes demonstrated nocturnal activity patterns, with coyotes more likely to be active during daylight hours. Overall, raccoons did not appear to exhibit high levels of vigilance. Activity pattern results provided further evidence that raccoons do not appear to fear coyotes, as both species were active at the same time and showed a high degree of overlap (Δ = 0.75) with little evidence of temporal segregation in activity. Thus, our study indicates that nonconsumptive effects of coyotes on raccoons are unlikely, which calls into question the ability of coyotes to initiate strong trophic cascades through some mesocarnivores.}, number={9}, journal={Diversity}, publisher={MDPI AG}, author={Chitwood, M. Colter and Lashley, Marcus A. and Higdon, Summer D. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2020}, month={Sep}, pages={341} } @article{boggs_moorman_hazel_greenberg_pacifici_2020, title={Relationships between white-footed mice and logging residue: Informing the sustainability of potential wood bioenergy harvests}, volume={457}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2019.117706}, abstractNote={We examined local and site-scale relationships between white-footed mice (Peromyscus leucopus) and logging residue after timber harvests to assess potential effects of expanding bioenergy markets in the southern Appalachian region of the United States. We sampled mice in 10 recent (2013–2015) clearcut or shelterwood harvests dominated either by white pine or hardwoods prior to harvest. We captured mice May–August, 2016 and 2017 using 10 grids of 60 Sherman traps spaced 15 m apart and set twice for five consecutive nights in each year. We categorized traps as either near (≤5 m) or far (>5 m) from coarse woody debris (CWD; woody debris ≥10 cm in diameter). We estimated site-level woody debris volumes using modified prism sweep sampling and determined vegetation, woody debris, and ground cover composition at each trap location. White-footed mouse occupancy increased with greater trap-level CWD cover in all stands, and greater site-level woody debris volume in white pine stands. Mouse abundance increased with greater site-level woody debris volume, and abundance was greater at white pine sites than hardwood sites. These results demonstrate that residual logging debris is important to white-footed mice, both at the local- and site-scale. Reductions in residual logging debris following harvests, including via removal of low value stems for wood bioenergy, likely will result in decreased white-footed mouse occupancy and abundance. We recommend developing proactive strategies to retain scattered logging residues following even-aged timber harvests, especially in cases where bioenergy harvests occur in the southern Appalachian region.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Boggs, April D. and Moorman, Christopher E. and Hazel, Dennis W. and Greenberg, Cathryn H. and Pacifici, Krishna}, year={2020}, month={Feb} } @article{michel_strickland_demarais_belant_kautz_duquette_beyer_chamberlain_miller_shuman_et al._2020, title={Relative reproductive phenology and synchrony affect neonate survival in a nonprecocial ungulate}, volume={34}, ISSN={0269-8463 1365-2435}, url={http://dx.doi.org/10.1111/1365-2435.13680}, DOI={10.1111/1365-2435.13680}, abstractNote={Abstract}, number={12}, journal={Functional Ecology}, publisher={Wiley}, author={Michel, Eric S. and Strickland, Bronson K. and Demarais, Stephen and Belant, Jerrold L. and Kautz, Todd M. and Duquette, Jared F. and Beyer, Dean E., Jr and Chamberlain, Michael J. and Miller, Karl V. and Shuman, Rebecca M. and et al.}, editor={Crocker, DanielEditor}, year={2020}, month={Sep}, pages={2536–2547} } @article{gefellers_buehler_moorman_zobel_harper_2020, title={Seeding is not always necessary to restore native early successional plant communities}, volume={28}, ISSN={["1526-100X"]}, DOI={10.1111/rec.13249}, abstractNote={Restoration of native early successional plant communities in the eastern United States is a conservation priority because of declining populations of associated plants and wildlife. Restoration typically involves seeding native species and is often fraught with problems including weedy competition, expensive seed, and slow establishment. Pairing seed bank response with strategic herbicide applications may be an alternative approach for restoring these plant communities. We compared early successional plant communities established by seeding (SD) paired with selective herbicide use to natural revegetation (NR) from the seed bank paired with selective herbicide use at 18 locations that were previously row‐crop or tall fescue (Schedonorus arundinaceus) fields in Tennessee, Alabama, and Kentucky, the United States. We did not detect differences in species diversity and richness, coverage of non‐native grasses and forbs, or number and coverage of native flowering forbs by season between NR and SD treatments at tall fescue or fallow crop sites. Species evenness was greatest in NR and coverage of native‐warm‐season grasses in SD. Species richness and coverage of native forbs were least in untreated tall fescue units (CNTL). More flexibility to use herbicides with NR reduced coverage of sericea lespedeza (Lespedeza cuneata) in NR units compared to SD units at tall fescue sites. NR was 3.7 times cheaper than seeding. Land managers should consider using an NR approach to establish native early successional plant communities.}, number={6}, journal={RESTORATION ECOLOGY}, author={GeFellers, James Wade and Buehler, David A. and Moorman, Christopher E. and Zobel, John M. and Harper, Craig A.}, year={2020}, month={Nov}, pages={1485–1494} } @article{kroeger_moorman_lashley_chitwood_harper_deperno_2020, title={White-tailed deer use of overstory hardwoods in longleaf pine woodlands}, volume={464}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/j.foreco.2020.118046}, DOI={10.1016/j.foreco.2020.118046}, abstractNote={Restoration of the longleaf pine ecosystem is a conservation priority throughout the southeastern United States, but the role of hardwoods in providing food and cover for wildlife within this system is poorly understood. We investigated white-tailed deer (Odocoileus virginianus) movement and habitat selection relative to overstory hardwood distribution in a longleaf pine ecosystem at Fort Bragg Military Installation in the Sandhills physiographic region of North Carolina from March 2011–July 2013. We monitored GPS-collared female white-tailed deer and used generalized linear mixed models and step-selection functions to determine the influence of overstory composition and understory cover on seasonal white-tailed deer habitat selection. During fall and winter, deer selection increased with increasing upland hardwood overstory until reaching an upper threshold (12% and 7%; respectively) where increasing cover of upland hardwoods no longer increased selection. Also, in the fall and winter, deer selected areas with greater bottomland hardwood overstory until an upper threshold of 33% bottomland hardwood overstory was reached. In the spring, deer selected areas with <22% upland hardwood overstory. The effect size of understory cover, defined as lidar-classified vegetation with height <2 m, was larger than any other variable, regardless of season, and deer consistently selected areas with 20–75% understory cover. When managing longleaf pine woodlands for white-tailed deer, our results indicate maintaining a well-developed woody understory with 20–50% canopy closure is important, ideally with mature upland hardwood overstory cover between 4 and 12% to ensure mast production in fall and winter.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Kroeger, Anthony J. and Moorman, Christopher E. and Lashley, Marcus A. and Chitwood, M. Colter and Harper, Craig A. and DePerno, Christopher S.}, year={2020}, month={May}, pages={118046} } @article{burke_peterson_sawyer_moorman_serenari_pacifici_2019, title={A method for mapping hunting occurrence using publicly available, geographic variables}, volume={43}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.994}, abstractNote={ABSTRACT}, number={3}, journal={WILDLIFE SOCIETY BULLETIN}, author={Burke, Conner R. and Peterson, M. Nils and Sawyer, David T. and Moorman, Christopher E. and Serenari, Christopher and Pacifici, Krishna}, year={2019}, month={Sep}, pages={537–545} } @article{taillie_moorman_smart_pacifici_2019, title={Bird community shifts associated with saltwater exposure in coastal forests at the leading edge of rising sea level}, volume={14}, ISSN={["1932-6203"]}, url={https://doi.org/10.1371/journal.pone.0216540}, DOI={10.1371/journal.pone.0216540}, abstractNote={Rising sea levels dramatically alter the vegetation composition and structure of coastal ecosystems. However, the implications of these changes for coastal wildlife are poorly understood. We aimed to quantify responses of avian communities to forest change (i.e., ghost forests) in a low-lying coastal region highly vulnerable to rising sea level. We conducted point counts to sample avian communities at 156 forested points in eastern North Carolina, USA in 2013–2015. We modelled avian community composition using a multi-species hierarchical occupancy model and used metrics of vegetation structure derived from Light Detection and Ranging (LiDAR) data as covariates related to variation in bird responses. We used this model to predict occupancy for each bird species in 2001 (using an analogous 2001 LiDAR dataset) and 2014 and used the change in occupancy probability to estimate habitat losses and gains at 3 spatial extents: 1) the entire study area, 2) burned forests only, and 3) unburned, low-lying coastal forests only. Of the 56 bird species we investigated, we observed parameter estimates corresponding to a higher likelihood of occurring in ghost forest for 34 species, but only 9 of those had 95% posterior intervals that did not overlap 0, thus having strong support. Despite the high vulnerability of forests in the region to sea level rise, habitat losses and gains associated with rising sea level were small relative to those resulting from wildfire. Though the extent of habitat changes associated with the development of ghost forest was limited, these changes likely are more permanent and may compound over time as sea level rises at an increasing rate. As such, the proliferation of ghost forests from rising sea level has potential to become an important driver of forest bird habitat change in coastal regions.}, number={5}, journal={PLOS ONE}, author={Taillie, Paul J. and Moorman, Christopher E. and Smart, Lindsey S. and Pacifici, Krishna}, year={2019}, month={May} } @article{bragina_kays_hody_moorman_deperno_mills_2019, title={Effects on white‐tailed deer following eastern coyote colonization}, volume={83}, ISSN={0022-541X 1937-2817}, url={http://dx.doi.org/10.1002/JWMG.21651}, DOI={10.1002/jwmg.21651}, abstractNote={ABSTRACT}, number={4}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Bragina, Eugenia V. and Kays, Roland and Hody, Allison and Moorman, Christopher E. and Deperno, Christopher S. and Mills, L. Scott}, year={2019}, month={Mar}, pages={916–924} } @article{drake_peterson_griffith_olfenbuttel_moorman_deperno_2019, title={Hunting interacts with socio-demographic predictors of human perceptions of urban coyotes}, volume={43}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.993}, abstractNote={ABSTRACT}, number={3}, journal={WILDLIFE SOCIETY BULLETIN}, author={Drake, Michael D. and Peterson, M. Nils and Griffith, Emily H. and Olfenbuttel, Colleen and Moorman, Christopher E. and Deperno, Christopher S.}, year={2019}, month={Sep}, pages={447–454} } @article{fish_moorman_schillaci_deperno_2019, title={Influence of military training on breeding ecology of Bachman's sparrow}, volume={83}, ISSN={0022-541X}, url={http://dx.doi.org/10.1002/jwmg.21579}, DOI={10.1002/jwmg.21579}, abstractNote={ABSTRACT}, number={1}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Fish, Alexander C. and Moorman, Christopher E. and Schillaci, Jessica M. and DePerno, Christopher S.}, year={2019}, month={Jan}, pages={72–79} } @article{taillie_moorman_2019, title={Marsh bird occupancy along the shoreline-to-forest gradient as marshes migrate from rising sea level}, volume={10}, ISSN={["2150-8925"]}, DOI={10.1002/ecs2.2555}, abstractNote={Abstract}, number={1}, journal={ECOSPHERE}, author={Taillie, Paul J. and Moorman, Christopher E.}, year={2019}, month={Jan} } @article{garabedian_moorman_peterson_kilgo_2019, title={Neighboring group density is more important than forest stand age to a threatened social woodpecker population}, ISSN={["1903-220X"]}, DOI={10.2981/wlb.00574}, abstractNote={Effective conservation of group-living forest wildlife requires information on how forest age moderates population parameters. Relationships between forest age and demographics can guide long-term management for wildlife populations that are expanding in relatively young second-growth forests in response to ongoing habitat management. We examined how forest age moderates effects of group density on long-term trends in group size and fledgling production in the endangered red-cockaded woodpecker Dryobates borealis (RCW) on the Savannah River Site, SC (SRS). We used 32 years of RCW monitoring data and generalized additive models to: 1) model long-term changes in average RCW group size and fledgling production; and 2) model effects of neighboring group density and neighboring group sizes across a gradient of forest age within 800 m of a group's cavity tree cluster. Average fledgling production oscillated over 2–3 year periods, but longer term evaluation indicated oscillations dampened and average fledgling production slightly decreased over time. Average group size fluctuated abruptly over 2–3 year periods from 1985 to 1994, but longer term evaluation indicated a general increase in group sizes from 1985 to 1994, followed by declines from 1995 to 2007, and a steady increase after 2010. Average fledgling production increased in response to neighboring group density but decreased as neighboring group sizes increased. In contrast, average group sizes increased in response to greater neighboring group density and neighboring group sizes. Stand age did not affect these relationships. Collectively, these results suggest forest age does not directly moderate effects of neighboring group density or group sizes on long-term average group size and fledgling production in the SRS RCW population. Although forest structure has been linked to increased RCW group sizes and productivity, our results suggest that with ongoing habitat management, long-term changes in group size and fledgling production will be driven primarily by group density conditions rather than changing forest age.}, journal={WILDLIFE BIOLOGY}, author={Garabedian, James E. and Moorman, Christopher E. and Peterson, M. Nils and Kilgo, John C.}, year={2019} } @article{rosche_moorman_pacifici_jones_deperno_2019, title={Northern bobwhite breeding season habitat selection in fire‐maintained pine woodland}, volume={83}, ISSN={0022-541X 1937-2817}, url={http://dx.doi.org/10.1002/jwmg.21683}, DOI={10.1002/jwmg.21683}, abstractNote={ABSTRACT}, number={5}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Rosche, Sarah B. and Moorman, ChristopheR E. and Pacifici, Krishna and Jones, Jeffrey G. and DePerno, Christopher S.}, year={2019}, month={May}, pages={1226–1236} } @book{moorman_grodsky_rupp_2019, place={Baltimore}, title={Renewable energy and wildlife conservation /}, publisher={Johns Hopkins University Press}, year={2019} } @article{stevenson_lashley_chitwood_garabedian_swingen_deperno_moorman_2019, title={Resource selection by coyotes (Canis latrans) in a longleaf pine (Pinus palustris) ecosystem: effects of anthropogenic fires and landscape features}, volume={97}, ISSN={0008-4301 1480-3283}, url={http://dx.doi.org/10.1139/cjz-2018-0150}, DOI={10.1139/cjz-2018-0150}, abstractNote={ Prescribed fire is used to restore and maintain fire-dependent forest communities. Because fire affects food and cover resources, fire-mediated resource selection has been documented for many wildlife species. The first step in understanding these interactions is to understand resource selection of the predators in a fire-maintained system. We attached GPS radio collars to 27 coyotes (Canis latrans Say, 1823) and examined resource selection relative to fire-maintained vegetation types, years since fire, anthropogenic features that facilitate prescribed burning, and other landscape features likely to affect coyote resource selection. Coyote home ranges were characterized by more open vegetation types and more recently burned forest (i.e., burned 0–1 year prior) than available on the study area. Within their home ranges, coyotes avoided areas close to densely vegetated drainages and paved roads. Coyote selection of more recently burned forest likely was in response to greater prey density or increased ability to detect prey soon after vegetation cover was reduced by fires; similarly, coyotes likely avoided drainages due to decreased hunting efficiency. Coyote resource selection was linked to prescribed fire, suggesting the interaction between fire and coyotes may influence ecosystem function in fire-dependent forests. }, number={2}, journal={Canadian Journal of Zoology}, publisher={Canadian Science Publishing}, author={Stevenson, E.R. and Lashley, M.A. and Chitwood, M.C. and Garabedian, J.E. and Swingen, M.B. and DePerno, C.S. and Moorman, C.E.}, year={2019}, month={Feb}, pages={165–171} } @article{sasmal_moorman_swingen_datta_deperno_2019, title={Seasonal space use of transient and resident coyotes (Canis latrans) in North Carolina, USA}, volume={97}, ISSN={["1480-3283"]}, DOI={10.1139/cjz-2018-0209}, abstractNote={ Coyote (Canis latrans Say, 1823) is a recent immigrant into eastern United States and little is known about the species’ space use and movement in the region. We compared space use and movement of radio-collared coyotes among biological seasons. We captured and collared 30 coyotes from February through May 2011 and collected 85 386 GPS locations through October 2012 at Fort Bragg Military Installation. We defined four biological seasons according to coyote life history: breeding (December–February), gestation (March–May), pup-rearing (June–August), and dispersal (September–November). Out of 27 radio-collared individuals, we identified 10 as transient and 11 as resident based on home-range size and variability across seasons; 6 switched their status and were classified as intermediate. We observed low variability of core-area size across seasons for resident males and females, whereas we documented high variability for transient males. Movement rate of resident coyotes during spring (449.75 m/h) was greater than summer (295.33 m/h), whereas movement rates did not differ between any other seasons. For transient coyotes, movement rate during summer (283 m/h) was less than fall (374.73 m/h), spring (479.85 m/h), and winter (488.5 m/h). Some coyotes adjusted their residency status seasonally and other individuals dispersed large distances (>200 km). }, number={4}, journal={CANADIAN JOURNAL OF ZOOLOGY}, author={Sasmal, Indrani and Moorman, Christopher E. and Swingen, Morgan B. and Datta, Shubham and DePerno, Christopher S.}, year={2019}, month={Apr}, pages={326–331} } @article{garabedian_peterson_moorman_kilgo_2019, title={Using qualitative methods to support recovery of endangered species: The case of red-cockaded woodpecker foraging habitat}, volume={17}, ISSN={2351-9894}, url={http://dx.doi.org/10.1016/J.GECCO.2019.E00553}, DOI={10.1016/j.gecco.2019.e00553}, abstractNote={Meta-analyses are powerful tools for synthesizing wildlife-habitat relationships, but small sample sizes and complex species-habitat relationships often preclude correlative meta-analyses on endangered species. In this study, we demonstrate qualitative comparative analysis (QCA) as a tool that can reliably synthesize habitat-fitness relationships from small sample sizes for species with narrow habitat requirements. We apply QCA to results from a habitat threshold regression tree model and identify habitat thresholds with consistent positive effects on fitness of the federally endangered red-cockaded woodpecker (Dryobates borealis; RCW) on the Savannah River Site, USA. We reformulated regression tree results in a QCA framework to examine the consistency of threshold effects on RCW fledgling production at the individual group level (n = 47). Synthesizing regression tree results with QCA revealed alternative combinations of habitat thresholds that in conjunction with group size consistently led to above-average fledgling production for 41 of 47 (88%) individual RCW groups. Importantly, QCA identified unique combinations of habitat thresholds and group size related to above-average fledgling production that were not retained in the regression tree model due to small sample sizes. Synthesizing a small habitat-fitness dataset using QCA provided a tractable method to identify unique combinations of habitat and group size conditions that are consistently important to individual fitness, but may not be detected by meta-analyses that can be biased by small sample sizes. QCA offers a viable approach for synthesis of habitat-fitness relationships and can be extended to address many complex issues in endangered species recovery when correlative meta-analyses are not possible.}, journal={Global Ecology and Conservation}, publisher={Elsevier BV}, author={Garabedian, James E. and Peterson, M. Nils and Moorman, Christopher E. and Kilgo, John C.}, year={2019}, month={Jan}, pages={e00553} } @article{bragina_kays_hody_moorman_deperno_mills_2019, title={White‐tailed deer and coyote colonization: a response to Kilgo et al. (2019)}, volume={83}, ISSN={0022-541X 1937-2817}, url={http://dx.doi.org/10.1002/jwmg.21766}, DOI={10.1002/jwmg.21766}, abstractNote={We stand by our conclusion that there have not been large‐scale declines in white‐tailed deer populations following coyote colonization of the eastern United States. However, we agree that coyote predation can affect deer populations locally and therefore and should be considered in deer harvest planning in the region.}, number={8}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Bragina, Eugenia V. and Kays, Roland and Hody, Allison and Moorman, Christopher E. and DePerno, Christopher S. and Mills, L. Scott}, year={2019}, month={Oct}, pages={1641–1643} } @article{chitwood_lashley_moorman_deperno_2018, title={Correction to: Setting an evolutionary trap: could the hider strategy be maladaptive for white-tailed deer?}, volume={36}, ISSN={0289-0771 1439-5444}, url={http://dx.doi.org/10.1007/S10164-017-0536-6}, DOI={10.1007/S10164-017-0536-6}, abstractNote={The article, “Setting an evolutionary trap: could the hider strategy be maladaptive for white-tailed deer?”, written by M. Colter Chitwood, Marcus A. Lashley, Christopher E. Moorman and Christopher S. DePerno, was originally published Online First without open access.}, number={2}, journal={Journal of Ethology}, publisher={Springer Nature}, author={Chitwood, M. Colter and Lashley, Marcus A. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2018}, month={May}, pages={215–215} } @article{sasmal_kilburg_deperno_chitwood_lashley_collier_moorman_2018, title={Eastern Wild Turkey Roost-site Selection in a Fire-maintained Longleaf Pine Ecosystem}, volume={17}, ISSN={["1938-5412"]}, DOI={10.1656/058.017.0301}, abstractNote={Abstract Night-time roosting in Meleagris gallopavo (Wild Turkey) is a quotidian activity that minimizes vulnerability to predators and weather. Roost-site selection in managed Pinus palustris (Longleaf Pine) communities is poorly documented. We assessed roost-site selection by comparing use and availability of vegetation types at the individual female Wild Turkey home-range level. We monitored 14 Wild Turkeys from February 2011 to June 2012. The Wild Turkeys did not use vegetation types within the estimated home ranges for roosting in proportion to availability (χ2 = 601.696, P < 0.001). Female Wild Turkeys roosted in the upland Longleaf Pine in proportion to availability, selected for lowland hardwood, and avoided upland hardwood patches. We documented that roost-site availability is not likely a limiting factor in managed Longleaf Pine forests.}, number={3}, journal={SOUTHEASTERN NATURALIST}, author={Sasmal, Indrani and Kilburg, Eric L. and DePerno, Christopher S. and Chitwood, M. Colter and Lashley, Marcus A. and Collier, Bret A. and Moorman, Christopher E.}, year={2018}, month={Sep}, pages={371–380} } @article{lashley_cove_chitwood_penido_gardner_deperno_moorman_2018, title={Estimating wildlife activity curves: comparison of methods and sample size}, volume={8}, ISSN={2045-2322}, url={http://dx.doi.org/10.1038/S41598-018-22638-6}, DOI={10.1038/S41598-018-22638-6}, abstractNote={Abstract}, number={1}, journal={Scientific Reports}, publisher={Springer Science and Business Media LLC}, author={Lashley, Marcus A. and Cove, Michael V. and Chitwood, M. Colter and Penido, Gabriel and Gardner, Beth and DePerno, Chris S. and Moorman, Chris E.}, year={2018}, month={Mar} } @article{garabedian_moorman_peterson_kilgo_2018, title={Evaluating interactions between space-use sharing and defence under increasing density conditions for the group-territorial Red-cockaded Woodpecker Leuconotopicus borealis}, volume={160}, ISSN={["1474-919X"]}, DOI={10.1111/ibi.12576}, abstractNote={Information about how bird species respond to increasing density conditions through either space‐use sharing or increased territoriality, and how those changes affect fitness, is essential for effective conservation planning. We used a case study of endangered Red‐cockaded Woodpeckers Leuconotopicus borealis (RCW) to address these questions. We documented over 36 000 locations from 44 RCW groups in three density conditions on two sites in South Carolina, USA, between April 2013 and March 2015. The frequency of neighbouring group interactions differed among density conditions and was highest for high‐density groups. RCW home‐ranges and core‐areas were larger under low‐density conditions ( = 88.4 ha,  = 21.0 ha) than under medium ( = 68.29 ha,  = 16.6 ha) and high‐density ( = 76.3 ha,  = 18.6 ha) conditions. Neighbouring RCWs maintained overlapping home‐ranges with nearly exclusive core‐areas across density conditions, but overlap tended to increase as neighbouring group density increased. Under high‐density conditions, home‐range overlap correlated inversely with clutch size (β ± se = −0.19 ± 0.09), nestling production (β ± se = −0.37 ± 0.09) and fledgling production (β ± se = −0.34 ± 0.08). Our results indicate that RCWs dedicate more effort to territorial defence under high‐density conditions, potentially at the expense of greater foraging efficiency and time allocated to reproduction, as evidenced by reduced fitness. Large home‐range overlap indicated limited territoriality farther away from cavity trees, but the existence of exclusive core‐areas suggests that RCW groups defend habitat closer to cavity trees. Thiessen partitions used to allocate critical foraging habitat offered comprehensive habitat protection for RCW but appear flawed for spatially explicit habitat assessments because they do not accurately delineate space used by individual RCW groups.}, number={4}, journal={IBIS}, author={Garabedian, James E. and Moorman, Christopher E. and Peterson, M. Nils and Kilgo, John C.}, year={2018}, month={Oct}, pages={816–831} } @article{taillie_burnett_roberts_campos_peterson_moorman_2018, title={Interacting and non-linear avian responses to mixed-severity wildfire and time since fire}, volume={9}, ISSN={2150-8925}, url={http://dx.doi.org/10.1002/ECS2.2291}, DOI={10.1002/ECS2.2291}, abstractNote={Abstract}, number={6}, journal={Ecosphere}, publisher={Wiley}, author={Taillie, Paul J. and Burnett, Ryan D. and Roberts, Lance Jay and Campos, Brent R. and Peterson, M. Nils and Moorman, Christopher E.}, year={2018}, month={Jun}, pages={e02291} } @article{grodsky_moorman_fritts_campbell_sorenson_bertone_castleberry_wigley_2018, title={Invertebrate community response to coarse woody debris removal for bioenergy production from intensively managed forests}, volume={28}, ISSN={["1939-5582"]}, DOI={10.1002/eap.1634}, abstractNote={Abstract}, number={1}, journal={ECOLOGICAL APPLICATIONS}, author={Grodsky, Steven M. and Moorman, Christopher E. and Fritts, Sarah R. and Campbell, Joshua W. and Sorenson, Clyde E. and Bertone, Matthew A. and Castleberry, Steven B. and Wigley, T. Bently}, year={2018}, month={Jan}, pages={135–148} } @article{greenberg_moorman_matthews-snoberger_waldrop_simon_heh_hagan_2018, title={Long-term herpetofaunal response to repeated fuel reduction treatments}, volume={82}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.21402}, abstractNote={ABSTRACT}, number={3}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Greenberg, Cathryn H. and Moorman, Christopher E. and Matthews-Snoberger, Charlotte E. and Waldrop, Thomas A. and Simon, Dean and Heh, Amanda and Hagan, Donald}, year={2018}, month={Apr}, pages={553–565} } @article{frew_peterson_sills_moorman_bondell_fuller_howell_2018, title={Market and Nonmarket Valuation of North Carolina's Tundra Swans among Hunters, Wildlife Watchers, and the Public}, volume={42}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.915}, abstractNote={ABSTRACT}, number={3}, journal={WILDLIFE SOCIETY BULLETIN}, author={Frew, Kristin N. and Peterson, M. Nils and Sills, Erin and Moorman, Christopher E. and Bondell, Howard and Fuller, Joseph C. and Howell, Douglas L.}, year={2018}, month={Sep}, pages={478–487} } @article{burke_peterson_sawyer_moorman_serenari_meentemeyer_deperno_2018, title={Predicting private landowner hunting access decisions and hunter density}, volume={24}, ISSN={1087-1209 1533-158X}, url={http://dx.doi.org/10.1080/10871209.2018.1545147}, DOI={10.1080/10871209.2018.1545147}, abstractNote={ABSTRACT Urbanization and shifting landowner demographics are changing how and where hunting occurs. We surveyed nonindustrial private landowners (N = 1,843) in North Carolina, USA to examine how demographics and land-use predict whether hunting occurred and hunter density. The optimal logistic regression model correctly predicted whether hunting occurred on 96% of properties. Larger properties, male property ownership, longer ownership tenure, income generation from a property, and landowners originating from rural environments were positively related to whether a property was hunted. Properties with older landowners and properties surrounded by greater housing and road density were less likely to be hunted. Hunter density declined with property size, longer ownership tenure, and the presence of a landowner or family member(s) hunting the property. In the future, increases in hunter density on small properties may facilitate wildlife management through hunting as landscapes become more urbanized.}, number={2}, journal={Human Dimensions of Wildlife}, publisher={Informa UK Limited}, author={Burke, Conner R. and Peterson, M. Nils and Sawyer, David T. and Moorman, Christopher E. and Serenari, Christopher and Meentemeyer, Ross K. and DePerno, Christopher S.}, year={2018}, month={Nov}, pages={99–115} } @article{fish_moorman_deperno_schillaci_hess_2018, title={Predictors of Bachman's Sparrow Occupancy at its Northern Range Limit}, volume={17}, ISSN={["1938-5412"]}, DOI={10.1656/058.017.0108}, abstractNote={Abstract Peucaea aestivalis (Bachman's Sparrow), a songbird endemic to the southeastern US, has experienced long-term population declines and a northern range-boundary retraction. Habitat loss and degradation, largely related to fire suppression, are believed to be the major causes of population declines, but these relationships are less studied at the northern range-extent. Hence, we investigated habitat selection of Bachman's Sparrow on Fort Bragg Military Installation, where vegetation is characterized by extensive fire-maintained Pinus palustris (Longleaf Pine) uplands. We surveyed breeding male sparrows using repeat-visit point-counts. We visited 182 points 3 times from April to July during the 2014 and 2015 breeding seasons. We measured vegetation and distance to other habitat features (e.g., wildlife openings, streams) at each point. We recorded presence or absence of Bachman's Sparrows and fit encounter histories into a single-season occupancy model in program Unmarked, including a year effect on detection. Occupancy probability was 0.52 and increased with greater grass-cover and at intermediate distances from wildlife openings, and decreased with years-since-fire and with greater shrub height. Predictors of Bachman's Sparrow occupancy were similar to those reported for other portions of the range, supporting the importance of frequent prescribed fire to maintain herbaceous groundcover used by birds for nesting and foraging. However, our study indicated that other habitat features (e.g., canopy openings) provided critical cover within extensive upland Longleaf Pine-Aristida stricta (Wiregrass) forest.}, number={1}, journal={SOUTHEASTERN NATURALIST}, author={Fish, Alexander C. and Moorman, Christopher E. and DePerno, Christopher S. and Schillaci, Jessica M. and Hess, George R.}, year={2018}, month={Mar}, pages={104–116} } @article{garabedian_moorman_peterson_kilgo_2018, title={Relative importance of social factors, conspecific density, and forest structure on space use by the endangered Red-cockaded Woodpecker: A new consideration for habitat restoration}, volume={120}, ISSN={["1938-5129"]}, DOI={10.1650/condor-17-211.1}, abstractNote={ABSTRACT Understanding how the interplay between social behaviors and habitat structure influences space use is important for conservation of birds in restored habitat. We integrated fine-grained LiDAR-derived habitat data, spatial distribution of cavity trees, and spatially explicit behavioral observations in a multi-scale model to determine the relative importance of conspecific density, intraspecific interactions, and the distribution of cavities on space use by Red-cockaded Woodpeckers (Picoides borealis) on 2 sites in South Carolina, USA. We evaluated candidate models using information theoretic methods. Top scale-specific models included effects of conspecific density and number of cavity tree starts within 200 m of Red-cockaded Woodpecker foraging locations, and effects of the number of intraspecific interactions within 400 m of Red-cockaded Woodpecker foraging locations. The top multi-scale model for 22 of 34 Red-cockaded Woodpecker groups included covariates for the number of groups within 200 m of foraging locations and LiDAR-derived habitat with moderate densities of large pines (Pinus spp.) and minimal hardwood overstory. These results indicate distribution of neighboring groups was the most important predictor of space use once a minimal set of structural habitat thresholds was reached, and that placing recruitment clusters as little as 400 m from foraging partitions of neighboring groups may promote establishment of new breeding groups in unoccupied habitat. The presence of neighboring groups likely provides cues to foraging Red-cockaded Woodpeckers that facilitate prospecting prior to juvenile dispersal and, to a lesser extent, indicates high-quality forage resources. Careful consideration of local distribution of neighboring groups in potential habitat may improve managers' ability to increase Red-cockaded Woodpecker density on restored landscapes and mitigate isolation of Red-cockaded Woodpecker groups, a problem that negatively affects fitness across the species' range.}, number={2}, journal={CONDOR}, author={Garabedian, James E. and Moorman, Christopher E. and Peterson, M. Nils and Kilgo, John C.}, year={2018}, month={May}, pages={305–318} } @article{greenberg_seiboldt_keyser_mcnab_scott_bush_moorman_2018, title={Reptile and amphibian response to season of burn in an upland hardwood forest}, volume={409}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2017.12.016}, abstractNote={Growing-season burns are increasingly used in upland hardwood forest for multiple forest management goals. Many species of reptiles and amphibians are ground-dwelling, potentially increasing their vulnerability to prescribed fire, especially during the growing-season when they are most active. We used drift fences with pitfall traps to experimentally assess how herpetofaunal species and communities responded to early, growing-season burns, dormant-season burns, and unburned controls. We documented no adverse effects of either growing-season burns or dormant-season burns on any common herpetofaunal taxa, but capture rates of total, adult, and juvenile five-lined skinks (Plestiodon fasciatus) were greater following growing-season burns. Most measurements reflected little or transient change in forest structure. However, canopy cover decreased by an average of 16% in growing-season burns within four growing-seasons of burning, with some tree mortality in patches where fire temperature likely was hotter. Our study suggests that even modest reductions in canopy cover may positively affect relative abundance and reproductive success of P. fasciatus. We cautiously suggest that a higher mean ground-level fire temperature and the physiologically active condition of vegetation in growing-season burns interacted to damage a greater proportion of trees, resulting in more canopy thinning than in dormant-season burns. However, weather, fuel types and condition, vegetation structure, and topography interact to affect fire intensity and the level of mortality or damage to canopy trees within and among stands, regardless of season conducted. We suggest that herpetofaunal response, for the species we studied, is more closely linked to change in canopy cover than to season of burn per se.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Greenberg, Cathryn H. and Seiboldt, Tyler and Keyser, Tara L. and McNab, W. Henry and Scott, Patrick and Bush, Janis and Moorman, Christopher E.}, year={2018}, month={Feb}, pages={808–816} } @article{bobay_taillie_moorman_2018, title={Use of autonomous recording units increased detection of a secretive marsh bird}, volume={89}, ISSN={["1557-9263"]}, DOI={10.1111/jofo.12274}, abstractNote={Obtaining sufficient numbers of detections during point counts to make inferences concerning the presence and abundance of secretive species, such as many species of marsh birds, can be difficult. However, autonomous recording units (ARUs) can provide extended survey windows, potentially allowing for more effective detection of elusive species. We assessed the feasibility of using both ARUs and point-count surveys to monitor Black Rails (Laterallus jamaicensis) and Least Bitterns (Ixobrychus exilis), two secretive marsh birds of conservation concern. We identified vocalizations in ARU recordings using acoustic analysis software, and combined these observations with those from point counts to model occupancy of both species in coastal marshes of eastern North Carolina in 2016 and 2017 while accounting for variation in detection. Use of ARUs doubled the number of points where we detected Black Rails; thus, the combined point count-ARU model yielded a greater occupancy probability for this species. However, the ARUs recorded few Least Bittern vocalizations, suggesting that successful application of ARUs may depend on the vocal complexity of focal species. Although the appropriateness of integrating ARUs with in-person monitoring varies among species, our results illustrate that this integration increased detections of an elusive species of conservation concern. RESUMEN. Uso de Unidades Aut onomas de Grabaci on incrementa la detecci on de un ave de pantano sigilosa La obtenci on de una cantidad suficiente de detecciones durante conteos por puntos, que permita hacer inferencias sobre la presencia y abundancia de especies sigilosas, como muchas especies de aves de pantano, puede ser dif ıcil. Sin embargo, el uso de unidades aut onomas de grabaci on (ARUs) puede ampliar estas oportunidades durante reconocimientos de campo, permitiendo potencialmente una detecci on m as efectiva de especies elusivas. Determinamos la factibilidad de uso de ARUs y reconocimientos por medio de conteos por puntos para registrar rascones (Laterallus jamaicensis) y avetoros (Ixobrychus exilis), dos aves de pantano sigilosas con estatus de conservaci on preocupante. Identificamos vocalizaciones en grabaciones con ARUs utilizando software de an alisis ac ustico y combinamos estas observaciones con aquellas de conteos por puntos para modelar la ocupaci on de ambas especies en los pantanos costeros del este de North Carolina en 2016 y 2017 mientras contabiliz abamos la variaci on de su detecci on. El uso de las ARUs duplic o el n umero de puntos donde detectamos a Laterallus jamaicensis, con lo que el modelo conteo por punto-ARU gener o una probabilidad de ocupaci on mayor para esta especie. Sin embargo, las ARUs grabaron pocas vocalizaciones de Ixobrychus exilis, lo que sugiere que el uso exitoso de ARUs podr ıa depender de la complejidad vocal de la especie focal. Aunque la pertinencia de integrar ARUs con registros hechos en persona var ıa entre especies, nuestros resultados ilustran que dicha integraci on incrementa las detecciones de una especie elusiva cuya conservaci on es preocupante.}, number={4}, journal={JOURNAL OF FIELD ORNITHOLOGY}, author={Bobay, Lucas R. and Taillie, Paul J. and Moorman, Christopher E.}, year={2018}, month={Dec}, pages={384–392} } @article{grodsky_campbell_fritts_wigley_moorman_2018, title={Variable responses of non-native and native ants to coarse woody debris removal following forest bioenergy harvests}, volume={427}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/J.FORECO.2018.02.010}, DOI={10.1016/J.FORECO.2018.02.010}, abstractNote={Timber harvests may facilitate ant invasions of forested landscapes, fostering interactions between non-native and native ants. Harvests that include removal of low-value woody biomass as forest bioenergy feedstock may reduce residual coarse woody debris, thereby altering food and cover resources for ant species. We manipulated: (1) volume and distribution of coarse woody debris in stand-scale treatments ranging from intensive coarse woody debris removal to no coarse woody debris removal; and (2) coarse woody debris availability at microsite locations within stand-scale treatments, including piles of hardwood stems, piles of conifer stems, and no pile locations in North Carolina, USA and windrows (i.e., long, linear piles of harvest residues) and no windrows in Georgia, USA, in recently clearcut pine plantations (n = 4 per state). We captured ants in regenerating stands and tested treatment- and location-level effects on non-native and native ant relative abundances. Invasive ants represented 19% of ant taxa richness, but comprised 94% of total ant captures. Red imported fire ant (Solenopsis invicta Buren, hereafter "RIFA") dominated the ant community in young plantations. RIFA avoided windrows, but its relative abundance did not differ among stand-scale treatments. Coarse woody debris retention in stand-scale treatments and at microsite locations favored non-RIFA ants, including Asian needle ant (Brachyponera chinensis Emery) and several native ant species. Dual invasions of RIFA and Asian needle ant in young plantations of the eastern United States may commonly occur because the two species may not compete for resources on the forest floor. Reduction of coarse woody debris via intensified woody biomass harvesting may negatively affect non-RIFA ant species and promote RIFA colonization, thereby indirectly increasing deleterious effects of RIFA on other wildlife.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Grodsky, Steven M. and Campbell, Joshua W. and Fritts, Sarah R. and Wigley, T. Bently and Moorman, Christopher E.}, year={2018}, month={Nov}, pages={414–422} } @article{mcalister_deperno_fuller_howell_moorman_2017, title={A Comparison of Field Methods to Estimate Canada Goose Abundance}, volume={41}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.827}, abstractNote={ABSTRACT}, number={4}, journal={WILDLIFE SOCIETY BULLETIN}, author={McAlister, Mark A. and DePerno, Christopher S. and Fuller, Joseph C. and Howell, Douglas L. and Moorman, Christopher E.}, year={2017}, month={Dec}, pages={685–690} } @article{chitwood_lashley_kilgo_cherry_conner_vukovich_ray_ruth_warren_deperno_et al._2017, title={Are camera surveys useful for assessing recruitment in white-tailed deer?}, volume={2017}, ISSN={0909-6396 1903-220X}, url={http://dx.doi.org/10.2981/wlb.00178}, DOI={10.2981/wlb.00178}, abstractNote={Camera surveys commonly are used by managers and hunters to estimate white‐tailed deer Odocoileus virginianus density and demographic rates. Though studies have documented biases and inaccuracies in the camera survey methodology, camera traps remain popular due to ease of use, cost‐effectiveness, and ability to survey large areas. Because recruitment is a key parameter in ungulate population dynamics, there is a growing need to test the effectiveness of camera surveys for assessing fawn recruitment. At Savannah River Site, South Carolina, we used six years of camera‐based recruitment estimates (i.e. fawn:doe ratio) to predict concurrently collected annual radiotag‐based survival estimates. The coefficient of determination (R2) was 0.445, indicating some support for the viability of cameras to reflect recruitment. We added two years of data from Fort Bragg Military Installation, North Carolina, which improved R2 to 0.621 without accounting for site‐specific variability. Also, we evaluated the correlation between year‐to‐year changes in recruitment and survival using the Savannah River Site data; R2 was 0.758, suggesting that camera‐based recruitment could be useful as an indicator of the trend in survival. Because so few researchers concurrently estimate survival and camera‐based recruitment, examining this relationship at larger spatial scales while controlling for numerous confounding variables remains difficult. Future research should test the validity of our results from other areas with varying deer and camera densities, as site (e.g. presence of feral pigs Sus scrofa) and demographic (e.g. fawn age at time of camera survey) parameters may have a large influence on detectability. Until such biases are fully quantified, we urge researchers and managers to use caution when advocating the use of camera‐based recruitment estimates.}, number={1}, journal={Wildlife Biology}, publisher={Wildlife Biology}, author={Chitwood, M. Colter and Lashley, Marcus A. and Kilgo, John C. and Cherry, Michael J. and Conner, L. Mike and Vukovich, Mark and Ray, H. Scott and Ruth, Charles and Warren, Robert J. and DePerno, Christopher S. and et al.}, year={2017}, month={Jan}, pages={wlb.00178} } @article{winiarski_moorman_carpenter_2017, title={Bachman's Sparrows at the northern periphery of their range: home range size and microhabitat selection}, volume={88}, ISSN={["1557-9263"]}, DOI={10.1111/jofo.12215}, abstractNote={Populations of Bachman's Sparrows (Peucaea aestivalis) have declined range-wide since the late 1960s. Populations at the periphery of their range have exhibited some of the steepest declines, and these sparrows are now rare or extirpated over much of the northern extent of their historical range. To better understand the spatial ecology of Bachman's Sparrows in this region of decline, we examined microhabitat selection and determined the home range sizes of radio-tagged male Bachman's Sparrows (N = 37) in the Coastal Plain of North Carolina in 2014 and 2015. From April to July, we located males 1–2 times daily for 5–6 d per week. We measured vegetation structure in home ranges using 5-m-radius plots centered on a subset of 10 randomly selected telemetry locations as well as in available unused locations 50 m and in a random direction from each telemetry location. Mean size of home ranges (7.9 ha) was larger than estimates reported in most previous studies, with differences among studies possibly due, at least in part, to differences in the characteristics of habitats where studies were conducted. The home ranges of Bachman's Sparrows in our study had greater densities of woody and dead vegetation than unused areas. Although generally considered detrimental to the presence of Bachman's Sparrows, the presence of some woody vegetation in frequently burned (i.e., ≤ 3-yr return interval) longleaf pine (Pinus palustris) communities like those in our study may be important in providing song perches for males and cover from attacking predators. Bachman's Sparrows in our study showed clear selection for several vegetation characteristics linked to frequent fire. Management strategies that approximate historical fire regimes in longleaf pine ecosystems should continue to be promoted as essential tools for the conservation of Bachman's Sparrows.}, number={3}, journal={JOURNAL OF FIELD ORNITHOLOGY}, author={Winiarski, Jason M. and Moorman, Christopher E. and Carpenter, John P.}, year={2017}, month={Sep}, pages={250–261} } @article{moorman_klimstra_harper_marcus_sorenson_2017, title={Breeding Songbird Use of Native Warm-Season and Non-Native Cool-Season Grass Forage Fields}, volume={41}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.726}, abstractNote={ABSTRACT}, number={1}, journal={WILDLIFE SOCIETY BULLETIN}, author={Moorman, Christopher E. and Klimstra, Ryan L. and Harper, Craig A. and Marcus, Jeffrey F. and Sorenson, Clyde E.}, year={2017}, month={Mar}, pages={42–48} } @article{chitwood_lashley_deperno_moorman_2017, title={Considerations on neonatal ungulate capture method: potential for bias in survival estimation and cause-specific mortality}, volume={2017}, ISSN={0909-6396 1903-220X}, url={http://dx.doi.org/10.2981/wlb.00250}, DOI={10.2981/wlb.00250}, abstractNote={A recent study of Sitka black‐tailed deer Odocoileus hemionus sitkensis demonstrated that opportunistic fawn capture yielded left‐truncated data and ultimately resulted in overestimating fawn survival and spurious ecological model inference compared to neonates captured via vaginal implant transmitters (VITs). Given the ecological and economic value of ungulates worldwide and the importance of neonate survival to understanding population dynamics, the potential biases in survival estimates and causes of mortality caused by left‐truncation must be transparent. Herein, we used a VIT‐based dataset from white‐tailed deer Odocoileus virginianus to examine potential problems with left‐truncated data. We manipulated our original VIT‐based dataset by randomly assigning age‐at‐capture to create three hypothetical opportunistic samples. We used the Kaplan—Meier estimator to quantify fawn survival to 16 weeks of age for the original and hypothetical datasets. Additionally, we compared the relative importance of mortality causes between the datasets. Survival for the original, VIT‐based dataset was 0.121 (SE = 0.043), while hypothetical datasets yielded overestimates (ranging from 0.191 to 0.234). The hypothetical opportunistic samples overestimated coyote predation as a source of mortality, while underestimating starvation. Because management actions rely on accurate estimates of survival and causes of mortality, we recommend that neonatal survival studies consider biases caused by capture method. For robust estimates of survival, VIT‐based samples appear to provide better estimates of survival, as opportunistic samples are biased high. We encourage future work to elucidate the potential for neonate capture technique to affect cause‐specific mortality.}, number={1}, journal={Wildlife Biology}, publisher={Wildlife Biology}, author={Chitwood, M. Colter and Lashley, Marcus A. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2017}, month={Jan}, pages={wlb.00250} } @article{lashley_chitwood_deperno_moorman_2017, title={Frequent fires eliminate fleshy fruit production}, volume={405}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2017.09.034}, abstractNote={Frequent fire-return intervals (<3-yr) have been suggested to optimize the benefits of prescribed fire in many fire-dominated ecosystems. There are several potential ecological benefits to frequent fires, such as suppression of encroaching fire-intolerant plant species, increased reproductive allocations of native herbaceous plant species, and increased plant diversity at the stand level. However, recent literature has reported a decline in frugivorous wildlife species in frequently burned landscapes, raising concern for fire-regime effects on fruit production. Thus, an assessment of the effects fire frequency on fleshy fruit abundance is needed. In a replicated field experiment following 4 or more rotations of a 1-yr, 2-yr, and 3-yr fire-return interval, we measured fruit production each month of the growing season (i.e., May-September) in the critically threatened longleaf pine (Pinus palustris) ecosystem – an ecosystem where frequent fire intervals commonly are recommended. Compared to the 3-yr fire-return interval, cumulative understory fruit production was 99% less following a 1-yr or 2-yr fire-return interval. In fact, all of the fruit detected in 1-yr and 2-yr treatments were detected in patches of vegetation unburned by the previous fire. Additionally, no fruits were detected on any transect in the midstory and overstory strata. These results suggest that applying fire on <3-yr fire-return intervals across large land areas could have negative effects on soft mast-dependent wildlife species. Moreover, without a mosaic in fire-spread, even a 3-yr fire return interval may eliminate midstory and overstory fleshy fruit production over time. We recommend fire managers incorporate multiple fire-return intervals and firing techniques to capture the ecological benefits of variability in frequency and spatial extents in fire.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Lashley, Marcus A. and Chitwood, M. Colter and DePerno, Christopher S. and Moorman, Christopher E.}, year={2017}, month={Dec}, pages={9–12} } @article{sasmal_deperno_swingen_moorman_2017, title={Influence of Vegetation Type and Prescribed Fire on Peromyscus Abundance in a Longleaf Pine Ecosystem}, volume={41}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.740}, abstractNote={ABSTRACT}, number={1}, journal={WILDLIFE SOCIETY BULLETIN}, author={Sasmal, Indrani and DePerno, Christopher S. and Swingen, Morgan B. and Moorman, Christopher E.}, year={2017}, month={Mar}, pages={49–54} } @article{chitwood_lashley_sherrill_sorenson_deperno_moorman_2017, title={Macroarthropod response to time-since-fire in the longleaf pine ecosystem}, volume={391}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2017.02.038}, abstractNote={Fire is an important disturbance worldwide, and literature supports the use of prescribed fire to restore and maintain fire-dependent ecosystems. However, fire could alter the abundance and persistence of some arthropods, in turn influencing vertebrate taxa that depend on those arthropods as a food source. We used replicated prescribed fire treatments to evaluate macroarthropod response to time-since-fire in the fire-maintained longleaf pine (Pinus palustris) ecosystem. We sampled macroarthropod assemblages using vinyl gutter pitfall traps for 5 consecutive days in each month of the study (May-August 2014) in each replicate burn block. We identified macroarthropods to Order and dried and weighed the samples to determine biomass (g) of all taxa detected. We focused our analyses on 4 macroarthropod taxa important as food for wild turkey (Meleagris gallopavo): Araneae, Coleoptera, Hymenoptera, and Orthoptera. We used standard least squares regression to evaluate the effect of time-since-fire on total biomass of the 4 Orders (and we also evaluated those Orders independently). The analysis indicated that time-since-fire had no effect (p = 0.2616) on combined biomass of these 4 taxa. Analyzing the 4 Orders separately, biomass of Araneae (p = 0.0057) and Orthoptera (p = 0.0004) showed significant effects of time-since-fire, while Coleoptera (p = 0.9465) and Hymenoptera (p = 0.1175) did not. Parameter estimates (Araneae = 0.0084; SE = 0.0029; Orthoptera = 0.0137; SE = 0.0036) indicated that greater time-since-fire resulted in greater biomass for those 2 Orders. Overall, time-since-fire did not appear to have substantial effects on macroarthropod biomass. However, responses by Araneae and Orthoptera provided evidence that longer time-since-fire may result in greatest levels of biomass for some taxa. Our results indicate the use of frequent prescribed fire to restore and maintain longleaf forests is unlikely to pose risks to overall macroarthropod biomass, particularly if heterogeneity in fire frequency and spatial extent occurs on the landscape.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Chitwood, M. Colter and Lashley, Marcus A. and Sherrill, Brandon L. and Sorenson, Clyde and DePerno, Christopher S. and Moorman, Christopher E.}, year={2017}, month={May}, pages={390–395} } @article{winiarski_fish_moorman_carpenter_deperno_schillaci_2017, title={Nest-site selection and nest survival of Bachman's Sparrows in two longleaf pine communities}, volume={119}, ISSN={["1938-5129"]}, DOI={10.1650/condor-16-220.1}, abstractNote={ABSTRACT Longleaf pine (Pinus palustris) ecosystems of the southeastern United States have experienced high rates of habitat loss and fragmentation, coinciding with dramatic population declines of a variety of taxa that inhabit the system. The Bachman's Sparrow (Peucaea aestivalis), a species closely associated with fire-maintained longleaf pine communities, is listed as a species of conservation concern across its entire range. Bachman's Sparrow breeding biology may provide valuable insights into population declines and inform restoration and management of remnant longleaf pine forest, but the species' secretive nesting habits have received little attention. We located 132 Bachman's Sparrow nests in the Coastal Plain and Sandhills physiographic regions of North Carolina, USA, during 2014–2015, and modeled nest-site selection and nest survival as a function of vegetation characteristics, burn history, temporal factors, and landscape-level habitat amount. There were distinct differences in nest-site selection between regions, with Bachman's Sparrows in the Coastal Plain region selecting greater woody vegetation density and lower grass density at nest sites than at non-nest locations. In contrast, sparrows selected nest sites with intermediate grass density and higher tree basal area in the Sandhills region. Despite clear patterns of nest-site selection, we detected no predictors of nest survival in the Sandhills, and nest survival varied only with date in the Coastal Plain. Daily survival rates were similar between regions, and were consistent with published studies from the species' core range where declines are less severe. Overall, our results indicate that creating and maintaining community-specific vegetation characteristics through the application of frequent prescribed fire should increase the amount of nesting cover for Bachman's Sparrows.}, number={3}, journal={CONDOR}, author={Winiarski, Jason M. and Fish, Alexander C. and Moorman, Christopher E. and Carpenter, John P. and DePerno, Christopher S. and Schillaci, Jessica M.}, year={2017}, month={Aug}, pages={361–374} } @article{lashley_chitwood_nanney_deperno_moorman_2017, title={Regenerating white pine (Pinus strobus) in the south: Seedling position is more important than herbivory protection}, volume={82}, DOI={10.2179/17-138}, abstractNote={ABSTRACT  Seedling survival and growth in eastern white pines (Pinus strobus L.) might be limited by white-tailed deer (Odocoileus virginianus) browsing. However, most studies have occurred in areas central to the white pine range, making other factors such as seedling microenvironment unimportant. If microenvironment becomes a concern near the edge of the white pine range, then factors such as seedling placement in relation to forest openings could be important, especially given that deer herbivory tends to be most intense near forest edges. We evaluated the relative importance of deer browse and seedling position in openings on seedling survival and growth in central North Carolina at the southern edge of the white pine range. Further, we determined if bud caps and caging improved survival and growth. Seedlings ≤ 10 m from the edge survived at a greater proportion than those > 10 m from the edge (83% and 73%, respectively). Initial height was the most important predictor of survival (R2 = 0.55; p < 0.01). When controlling for initial seedling height, the location of the seedling (p < 0.01) within the opening was the only significant predictor of survival, despite the increase of browse near the edges of openings on unprotected seedlings. Caging and bud caps decreased seedling browse by 80% but had no effect on subsequent seedling survival (p = 0.28). A smaller proportion of seedlings with bud caps survived—an effect exacerbated by being internal to the opening. Our data indicate seedling microenvironment is an important consideration at the periphery of the white pine range.}, number={2}, journal={Castanea}, author={Lashley, M. A. and Chitwood, M. C. and Nanney, J. S. and DePerno, Chris and Moorman, C. E.}, year={2017}, pages={156–162} } @article{winiarski_moorman_carpenter_hess_2017, title={Reproductive consequences of habitat fragmentation for a declining resident bird of the longleaf pine ecosystem}, volume={8}, ISSN={2150-8925}, url={http://dx.doi.org/10.1002/ECS2.1898}, DOI={10.1002/ECS2.1898}, abstractNote={Abstract}, number={7}, journal={Ecosphere}, publisher={Wiley}, author={Winiarski, Jason M. and Moorman, Christopher E. and Carpenter, John P. and Hess, George R.}, year={2017}, month={Jul}, pages={e01898} } @article{fritts_moorman_grodsky_hazel_homyack_farrell_castleberry_evans_greene_2017, title={Rodent response to harvesting woody biomass for bioenergy production}, volume={81}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.21301}, abstractNote={ABSTRACT}, number={7}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Fritts, Sarah R. and Moorman, Christopher E. and Grodsky, Steven M. and Hazel, Dennis W. and Homyack, Jessica A. and Farrell, Christopher B. and Castleberry, Steven B. and Evans, Emily H. and Greene, Daniel U.}, year={2017}, month={Sep}, pages={1170–1178} } @article{chitwood_lashley_moorman_deperno_2017, title={Setting an evolutionary trap: could the hider strategy be maladaptive for white-tailed deer?}, volume={35}, ISSN={["1439-5444"]}, DOI={10.1007/s10164-017-0514-z}, abstractNote={Abstract}, number={3}, journal={JOURNAL OF ETHOLOGY}, author={Chitwood, M. Colter and Lashley, Marcus A. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2017}, month={Sep}, pages={251–257} } @article{biggerstaff_lashley_chitwood_moorman_deperno_2017, title={Sexual segregation of forage patch use: Support for the social-factors and predation hypotheses}, volume={136}, ISSN={["1872-8308"]}, DOI={10.1016/j.beproc.2017.01.003}, abstractNote={Nearly all species of sexually dimorphic ungulates sexually segregate. Several hypotheses have been proposed to explain this phenomenon, including the social-factors hypothesis (SFH) and the predation hypothesis (PH). Interestingly, previous studies have accepted and rejected each hypothesis within and across species but few studies have simultaneously tested both hypotheses in the same population. In August 2011 and 2012 using 7680 photographs taken with camera traps in standardized forage patches, we tested two predictions of the SFH: 1) foraging efficiency of both sexes would decrease when foraging rate in mixed-sex groups relative to single-sex groups, and 2) activity patterns (i.e., the pattern of temporal use of forage patches on a diel scale) of the sexes would decrease in temporal overlap at the forage patch level (i.e., social segregation) compared to the overall temporal overlap of activity patterns of the population. Also, we tested two predictions of the +PH : 1) the relationship between feeding rates of each sex, and 2) temporal activity overlap would change with changing risk level of forage patches as a result of differing risk perception between sexes. In support of the SFH for temporal segregation, when in mixed-sex groups, mature males and all females decreased feeding rate 30% and 10%, respectively; further, the sexes had similar activity patterns overall (94-95% overlap), though temporal overlap was lower in individual forage patches (68-74% overlap). In multi-male mixed sex groups, at least one male exhibited aggressive posture toward females during all foraging bouts suggesting intersex aggression was the cause of the observed decrease in foraging rates. In support of the PH , the sexes adjusted feeding rate differently in response to changing risk level of a forage patch, encouraging spatial segregation; however, the PH was not supported for temporal segregation because temporal activity pattern overlap did not vary as a function of predation risk. Coupling our results with previous reports indicates that the SFH is supported for only temporal segregation of forage patch use, and the PH may only be supported for spatial segregation in forage patch use. Thus, both social factors and predation risk may interact to encourage sexual segregation.}, journal={BEHAVIOURAL PROCESSES}, author={Biggerstaff, Michael T. and Lashley, Marcus A. and Chitwood, M. Colter and Moorman, Christopher E. and DePerno, Christopher S.}, year={2017}, month={Mar}, pages={36–42} } @article{morina_lashley_chitwood_moorman_deperno_2017, title={Should We Use the Float Test to Quantify Acorn Viability?}, volume={41}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.826}, abstractNote={ABSTRACT}, number={4}, journal={WILDLIFE SOCIETY BULLETIN}, author={Morina, Daniel L. and Lashley, Marcus A. and Chitwood, M. Colter and Moorman, Christopher E. and DePerno, Christopher S.}, year={2017}, month={Dec}, pages={776–779} } @article{garabedian_moorman_peterson_kilgo_2017, title={Use of LiDAR to define habitat thresholds for forest bird conservation}, volume={399}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2017.05.024}, abstractNote={Quantifying species-habitat relationships provides guidance for establishment of recovery standards for endangered species, but research on forest bird habitat has been limited by availability of fine-grained forest structure data across broad extents. New tools for collection of data on forest bird response to fine-grained forest structure provide opportunities to evaluate habitat thresholds for forest birds. We used LiDAR-derived estimates of habitat attributes and resource selection to evaluate foraging habitat thresholds for recovery of the federally endangered red-cockaded woodpecker (Leuconotopicus borealis; RCW) on the Savannah River Site, South Carolina. First, we generated utilization distributions to define habitat use and availability for 30 RCW groups surveyed over a >4-h period twice per month between April 2013 and March 2015. Next, we used piecewise regression to characterize RCW threshold responses to LiDAR-derived habitat attributes described in the United States Fish and Wildlife Service recovery plan for RCW. Finally, we used resource utilization functions to estimate selection of specific habitat thresholds and used the magnitude of selection to prioritize thresholds for conservation. We identified lower and upper thresholds for densities of pines ≥35.6 cm dbh (22, 65 trees/ha), basal area (BA) of pines ≥25.4 cm dbh (1.4, 2.2 m2/ha), hardwood canopy cover (6, 31%), and BA of hardwoods 7.6–22.9 cm dbh (0.4, 6.07 m2/ha); we identified three thresholds for density of pines 7.6–25.4 cm dbh (56, 341, and 401 trees/ha). Selection rankings prioritized foraging habitat with <6% hardwood canopy cover (β = 0.254, 95% CI = 0.172–0.336), < 1.2 m2/ha BA of hardwoods 7.6–22.9 cm dbh (β = 0.162, 95% CI = 0.050–0.275), ≥1.4 m2/ha BA of pines ≥25.4 cm dbh (β = 0.055, 95% CI = 0.022–0.087), and ≥22 pines ≥35.6 cm dbh/ha (β = 0.015, 95% CI = 0.013–0.042). We identified habitat thresholds corresponding to open canopy structure, moderate densities of large and medium pines, and sparse hardwood midstory trees. Selection ranks prioritized multiple thresholds below USFWS range-wide recovery thresholds, indicating site-specific management goals may be beneficial for RCW conservation. Fine-grained LiDAR-derived habitat data coupled with GPS-derived habitat use can guide forest bird conservation by identifying the full range of structural conditions associated with threshold responses.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Garabedian, James E. and Moorman, Christopher E. and Peterson, M. Nils and Kilgo, John C.}, year={2017}, month={Sep}, pages={24–36} } @article{mcalister_moorman_meentemeyer_fuller_howell_deperno_2017, title={Using Landscape Characteristics to Predict Distribution of Temperate-Breeding Canada Geese}, volume={16}, ISSN={["1938-5412"]}, DOI={10.1656/058.016.0201}, abstractNote={Abstract Accurate estimates of species' distributions are needed to ensure that conservation-planning efforts are directed at appropriate areas. Since the early 1980s, temperate-breeding populations of Branta canadensis (Canada Goose) have increased, yet reliable estimates of the species' distribution are lacking in many regions. Our objective was to identify the landcover features that best predicted Canada Goose distribution. In April 2015, we surveyed 300 one-km2 plots across North Carolina and observed 449 Canada Geese. We quantified percent coverage of 7 continuous landcover variables at 5 different spatial extents for each of the 300 plots. We fit logistic regression models using presence and absence at the 300 plots as the dependent variable and percent-cover covariates as independent variables. The best model for predicting Canada Goose presence included percent pasture within the 9 km2 surrounding the survey plot and percent open water within the 1-km2 survey plot. The probability of Canada Goose presence increased with increasing percent open water and percent pasture, albeit at different spatial extents, which provided important cover and food resources, respectively. Our approach using remote-sensing data to accurately predict Canada Goose presence across a large spatial extent can be employed to determine distributions for other easily surveyed, widely distributed species.}, number={2}, journal={SOUTHEASTERN NATURALIST}, author={McAlister, Mark A. and Moorman, Christopher E. and Meentemeyer, Ross K. and Fuller, Joseph C. and Howell, Douglas L. and DePerno, Christopher S.}, year={2017}, month={Jun}, pages={127–139} } @article{grodsky_moorman_fritts_castleberry_wigley_2016, title={Breeding, Early-Successional Bird Response to Forest Harvests for Bioenergy}, volume={11}, ISSN={["1932-6203"]}, DOI={10.1371/journal.pone.0165070}, abstractNote={Forest regeneration following timber harvest is a principal source of habitat for early-successional birds and characterized by influxes of early-successional vegetation and residual downed woody material. Early-successional birds may use harvest residues for communication, cover, foraging, and nesting. Yet, increased market viability of woody biomass as bioenergy feedstock may intensify harvest residue removal. Our objectives were to: 1) evaluate effects of varying intensities of woody biomass harvest on the early-successional bird community; and (2) document early-successional bird use of harvest residues in regenerating stands. We spot-mapped birds from 15 April– 15 July, 2012–2014, in six woody biomass removal treatments within regenerating stands in North Carolina (n = 4) and Georgia (n = 4), USA. Treatments included clearcut harvest followed by: (1) traditional woody biomass harvest with no specific retention target; (2) 15% retention with harvest residues dispersed; (3) 15% retention with harvest residues clustered; (4) 30% retention with harvest residues dispersed; (5) 30% retention with harvest residues clustered; and (6) no woody biomass harvest (i.e., reference site). We tested for treatment-level effects on breeding bird species diversity and richness, early-successional focal species territory density (combined and individual species), counts of breeding birds detected near, in, or on branches of harvest piles/windrows, counts of breeding bird behaviors, and vegetation composition and structure. Pooled across three breeding seasons, we delineated 536 and 654 territories and detected 2,489 and 4,204 birds in the North Carolina and Georgia treatments, respectively. Woody biomass harvest had limited or short-lived effects on the early-successional, breeding bird community. The successional trajectory of vegetation structure, rather than availability of harvest residues, primarily drove avian use of regenerating stands. However, many breeding bird species used downed wood in addition to vegetation, indicating that harvest residues initially may provide food and cover resources for early-successional birds in regenerating stands prior to vegetation regrowth.}, number={10}, journal={PLOS ONE}, author={Grodsky, Steven M. and Moorman, Christopher E. and Fritts, Sarah R. and Castleberry, Steven B. and Wigley, T. Bently}, year={2016}, month={Oct} } @article{bohling_dellinger_mcvey_cobb_moorman_waits_2016, title={Describing a developing hybrid zone between red wolves and coyotes in eastern North Carolina, USA}, volume={9}, ISSN={["1752-4571"]}, DOI={10.1111/eva.12388}, abstractNote={Abstract}, number={6}, journal={EVOLUTIONARY APPLICATIONS}, author={Bohling, Justin H. and Dellinger, Justin and McVey, Justin M. and Cobb, David T. and Moorman, Christopher E. and Waits, Lisette P.}, year={2016}, month={Jul}, pages={791–804} } @article{fritts_moorman_grodsky_hazel_homyack_farrell_castleberry_2016, title={Do biomass harvesting guidelines influence herpetofauna following harvests of logging residues for renewable energy?}, volume={26}, ISSN={["1939-5582"]}, DOI={10.1890/14-2078}, abstractNote={Abstract}, number={3}, journal={ECOLOGICAL APPLICATIONS}, author={Fritts, Sarah and Moorman, Christopher and Grodsky, Steven and Hazel, Dennis and Homyack, Jessica and Farrell, Chris and Castleberry, Steven}, year={2016}, month={Apr}, pages={926–939} } @article{lashley_chitwood_street_moorman_deperno_2016, title={Do indirect bite count surveys accurately represent diet selection of white-tailed deer in a forested environment?}, volume={43}, ISSN={["1448-5494"]}, DOI={10.1071/wr15008}, abstractNote={ Context Diet selection is studied in herbivores using three predominant methods: (1) microhistological surveys (identification of plants cell walls remaining in gut contents or faecal excretions); (2) direct bite counts (of tame animals); and (3) indirect bite counts (identifying herbivory on damaged plant tissues). Microhistological surveys and direct bite counts are accurate and provide the potential advantage of linking diet selection to particular individuals. Also, they allow diet selection to be measured in systems with sympatric herbivores more easily than indirect bite counts. However, they require expertise in cell wall structure identification or access to tame animals, and generally require greater expense than indirect bite counts. Conversely, indirect bite counts have the advantages of relatively low cost and time commitment for gathering data and do not require animal observation, but may not be accurate. Aims We tested for similarity between diet-selection estimates calculated by indirect bite counts and microhistological surveys. Methods We performed concurrent indirect bite count and faecal microhistological surveys on white-tailed deer (Odocoileus virginianus) at Fort Bragg Military Installation, NC. Key results The indirect bite count survey assignment of selection was 48% similar to assignments derived from the microhistological survey, based on Jaccard’s similarity index. Out of 23 plant species determined to be selected by indirect bite counts, 15 of those species were selected according to microhistological surveys. According to the microhistological survey, eight of the selected plants made up 51% of the overall diet, and seven of those eight were selected according to the indirect bite counts. Conclusions Our data indicate that indirect bite counts may provide a relatively accurate index of the deer-selected plants most important in the white-tailed deer diet, but may be less appropriate to determine selection of plants that infrequently occur in their diet, plants that are typically consumed in entirety, or plants where herbivory damage is poorly identified. Implications Indirect bite counts are a relatively inexpensive and time-efficient tool that may be useful to determine plant species most important to white-tailed deer within a forested landscape, particularly if additional research can improve on associated inaccuracies. }, number={3}, journal={WILDLIFE RESEARCH}, author={Lashley, Marcus A. and Chitwood, M. Colter and Street, Garrett M. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2016}, pages={254–260} } @article{rodriguez_peterson_moorman_2016, title={Does education influence wildlife friendly landscaping preferences?}, volume={20}, ISSN={1083-8155 1573-1642}, url={http://dx.doi.org/10.1007/S11252-016-0609-2}, DOI={10.1007/s11252-016-0609-2}, number={2}, journal={Urban Ecosystems}, publisher={Springer Nature}, author={Rodriguez, Shari L. and Peterson, M. Nils and Moorman, Christopher J.}, year={2016}, month={Nov}, pages={489–496} } @misc{harper_ford_lashley_moorman_stambaugh_2016, title={FIRE EFFECTS ON WILDLIFE IN THE CENTRAL HARDWOODS AND APPALACHIAN REGIONS, USA}, volume={12}, ISSN={["1933-9747"]}, DOI={10.4996/fireecology.1202127}, abstractNote={Fire is being prescribed and used increasingly to promote ecosystem restoration (e.g., oak woodlands and savannas) and to manage wildlife habitat in the Central Hardwoods and Appalachian regions, USA. However, questions persist as to how fire affects hardwood forest communities and associated wildlife, and how fire should be used to achieve management goals. We provide an up-to-date review of fire effects on various wildlife species and their habitat in the Central Hardwoods and Appalachians. Documented direct effects (i.e., mortality) on wildlife are rare. Indirect effects (i.e., changes in habitat quality) are influenced greatly by light availability, fire frequency, and fire intensity. Unless fire intensity is great enough to kill a portion of the overstory, burning in closed-canopy forests has provided little benefit for most wildlife species in the region because it doesn’t result in enough sunlight penetration to elicit understory response. Canopy reduction through silvicultural treatment has enabled managers to use fire more effectively. Fire intensity must be kept low in hardwoods to limit damage to many species of overstory trees. However, wounding or killing trees with fire benefits many wildlife species by allowing increased sunlight to stimulate understory response, snag and subsequent cavity creation, and additions of large coarse woody debris. In general, a fire-return interval of 2 yr to 7 yr benefits a wide variety of wildlife species by providing a diverse structure in the understory; increasing browse, forage, and soft mast; and creating snags and cavities. Historically, dormant-season fire was most prevalent in these regions, and it still is when most prescribed fire is implemented in hardwood systems as burn-days are relatively few in the growing season of May through August because of shading from leaf cover and high fuel moisture. Late growing-season burning increases the window for burning, and better control on woody composition is possible. Early growing-season fire may pose increased risk for some species, especially herpetofauna recently emerged from winter hibernacula (April) or forest songbirds that nest in the understory (May to June). However, negative population-level effects are unlikely unless the burned area is relatively large and early growing-season fire is used continually. We did not find evidence that fire is leading to population declines for any species, including Endangered Species Act (ESA)-listed species (e.g., Indiana bat [Myotis sodalis Mill. Allen] or northern long-eared bat [M. septentrionalis Trouess.]). Instead, data indicate that fire can enhance habitat for bats by increasing suitability of foraging and day-roost sites. Similarly, concern over burning and displacement of woodland salamanders (Plethodontidae), another taxa of heightened conservation concern, is alleviated when fire is prescribed along ecologically appropriate aspect and slope gradients and not forced into mesic, high site index environments where salamanders are most common. Because topography across the Central Hardwoods and Appalachians is diverse, we contend that applying fire on positions best suited for burning is an effective approach to increase regional landscape heterogeneity and biological diversity. Herein, we offer prescriptive concepts for burning for various wildlife species and guilds in the Central Hardwoods and Appalachians.ResumenEl fuego prescrito está siendo propuesto y utilizado cada vez más para promover la restauración de los ecosistemas (por ej. los arbustales de robles y sabanas) y para manejar el hábitat de la fauna silvestre en los Bosques Centrales de Latifoliadas y en los Apalaches en EEUU. Sin embargo, persiste el interrogante en como el fuego afecta a las comunidades del bosque de latifoliadas y su fauna asociada, y cómo el fuego debería ser utilizado para lograr objetivos de manejo. Nosotros realizamos una revisión actualizada de los efectos del fuego en varias especies de fauna silvestre y su hábitat en los Bosques Centrales de Latifoliadas y los Apalaches. La documentación sobre los efectos directos del fuego (por ej., mortalidad) en la fauna silvestre son raros. Los efectos indirectos (por ej., cambios en la calidad del hábitat) son influenciados grandemente por la disponibilidad de luz, y la frecuencia y la intensidad del fuego. A menos que la intensidad del fuego sea lo suficientemente grande como para matar una porción del estrato superior, las quemas en bosques con el canopeo cerrado han aportado poco beneficio a la mayoría de las especies de fauna silvestre en la región, porque no permiten la penetración de suficiente de luz solar para obtener una respuesta en el sotobosque. La reducción del canopeo a través de tratamientos silviculturales ha permitido a los gestores utilizar el fuego en forma más efectiva. La intensidad del fuego debe ser mantenida baja en bosques de latifoliadas para limitar el daño a algunas especies arbóreas del estrato superior. Sin embargo, hiriendo o matando árboles mediante su quema puede beneficiar a muchas especies de la fauna silvestre, al permitir así la entrada de la luz solar para estimular la respuesta del sotobosque, la creación de árboles muertos en pie y la subsecuente formación de cavidades en ellos, y la adición de restos leñosos gruesos. En general, un intervalo de retorno del fuego de 2 años a 7 años beneficia a una amplia variedad de especies de la fauna silvestre, proporcionando una estructura diversa en el sotobosque, incrementando el ramoneo, el forraje, y los frutos carnosos, y creando troncos muertos en pié y cavidades en ellos. Históricamente, el fuego en la temporada de dormancia era más preponderante en estas regiones, y todavía lo es cuando la mayoría de los fuegos prescriptos se implementan en sistemas de latifoliadas, dado que los días de quema son relativamente pocos en la temporada de crecimiento de mayo hasta agosto debido a la sombra de la cobertura de hojas y el alto contenido de humedad de los combustibles. Las quemas durante la temporada tardía de crecimiento incrementan la ventana de prescripción, y es posible un mejor control en la composición de leñosas. Los fuegos al principio de la temporada de crecimiento pueden ocasionar mayores riesgos para algunas especies, especialmente para la herpetofauna recién emergida de la hibernación (abril) o a los pájaros cantores del bosque que anidan en el sotobosque (mayo a junio). Sin embargo, efectos negativos a nivel de población son improbables a menos que el área quemada sea relativamente grande y las quemas sean utilizadas en la temporada temprana de crecimiento en forma continua. Nosotros no hemos encontrado evidencias de que el fuego esté conduciendo a la declinación de la población de cualquiera de las especies, incluyendo las listadas en la Ley de Especies Amenazadas (ESA) (por ej., el murciélago de Indiana [Myotsis sodalis Mill. Allen] o el murciélago orejudo [M. septentrionalis Touess.]). En cambio, los datos indican que el fuego puede favorecer el hábitat de los murciélagos al aumentar la capacidad de alimentarse y los sitios de percheo diurnos. En forma similar, una preocupación sobre las quemas y el desplazamiento de la salamandra de los arbustales (Plethodontidae), otro de los taxones de alto interés de conservación, se alivia cuando el fuego se prescribe a lo largo de orientaciones y gradientes de pendiente apropiados ecológicamente y no forzados a ambientes mésicos, donde los índices de presencia en el ambiente son altos y las salamandras son más comunes. Debido a que la topografía a través de los Bosques Centrales de Latifoliadas y de los Apalaches es diversa, nosotros bregamos para que la aplicación de fuego en las posiciones más apropiadas sea un acercamiento efectivo para aumentar la heterogeneidad regional del paisaje y la diversidad biológica. Al respecto, nosotros ofrecemos conceptos de prescripción para las quemas de varias especies de la fauna silvestre y especies asociadas en los Bosques Centrales de Latifoliadas y los Apalaches.}, number={2}, journal={FIRE ECOLOGY}, author={Harper, Craig A. and Ford, W. Mark and Lashley, Marcus A. and Moorman, Christopher E. and Stambaugh, Michael C.}, year={2016}, pages={127–159} } @article{gardner_garner_cobb_moorman_2016, title={Factors Affecting Occupancy and Abundance of American Alligators at the Northern Extent of Their Range}, volume={50}, ISSN={["1937-2418"]}, DOI={10.1670/15-147}, abstractNote={Abstract Populations of American Alligators (Alligator mississippiensis) generally are considered more abundant at present than historically; however, little information exists to assess the population of alligators in North Carolina at the northern extent of the species' range. Investigation of the factors influencing the distribution and abundance of alligators in North Carolina could shed light on the species' response to rapid environmental change in the region. We conducted a two-phase study: 1) to assess the distribution of alligators in North Carolina using a site-occupancy design; and 2) to assess the patterns in abundance using a repeated sampling design for population estimation. Results showed that both occupancy and abundance decreased in more northern sites, in sites with higher salinity, and in sites that were generally more westward. Sites sampled later in June were more likely to be occupied than those sampled earlier in the month. Abundance also increased with greater shoreline vegetation complexity and varied between lakes, rivers, and estuaries. Compared with studies from 30 years prior, the population seems fairly stable in terms of abundance and distribution. Given the northern limits of the species and the negative association with salinity, continued monitoring is warranted to understand changes in distribution and abundance with respect to predicted rates of sea-level rise, salinization, and urbanization locally around coastal cities like Wilmington.}, number={4}, journal={JOURNAL OF HERPETOLOGY}, author={Gardner, Beth and Garner, Lindsey A. and Cobb, David T. and Moorman, Christopher E.}, year={2016}, month={Dec}, pages={541–547} } @article{greenberg_moorman_raybuck_sundol_keyser_bush_simon_warburton_2016, title={Reptile and amphibian response to oak regeneration treatments in productive southern Appalachian hardwood forest}, volume={377}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2016.06.023}, abstractNote={Forest restoration efforts commonly employ silvicultural methods that alter light and competition to influence species composition. Changes to forest structure and microclimate may adversely affect some taxa (e.g., terrestrial salamanders), but positively affect others (e.g., early successional birds). Salamanders are cited as indicators of ecosystem health because of their sensitivity to forest floor microclimate. We used drift fences with pitfall and funnel traps in a replicated Before-After-Control-Impact design to experimentally assess herpetofaunal community response to initial application of three silvicultural methods proposed to promote oak regeneration: prescribed burning; midstory herbicide; and shelterwood harvests (initial treatment of the shelterwood-burn method) and controls, before and for five years post-treatment. Species richness of all herpetofauna, amphibians, reptiles, frogs, salamanders, or snakes was unaffected by any treatment, but lizard species richness increased in the shelterwood harvest. Capture rate of total salamanders decreased post-harvest in shelterwood units after a 2–3 year delay; Plethodon teyahalee decreased post-harvest in shelterwoods, but also in control units. In contrast, capture rate of total lizards and Plestiodon fasciatus increased in shelterwood stands within the first year post-harvest. Prescribed burn and midstory herbicide treatments did not affect any reptile or amphibian species. A marginally lower proportion of juvenile to adult P. teyahalee, and a higher proportion of juvenile P. fasciatus in shelterwood than control units suggested that heavy canopy removal and associated change in microclimate may differentially affect reproductive success among species. Our study illustrates the importance of longer-term studies to detect potential changes in herpetofaunal communities that may not be immediately apparent after disturbances, and highlights the importance of including multiple taxa for a balanced perspective when weighing impacts of forest management activities.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Greenberg, Cathryn H. and Moorman, Christopher E. and Raybuck, Amy L. and Sundol, Chad and Keyser, Tara L. and Bush, Janis and Simon, Dean M. and Warburton, Gordon S.}, year={2016}, month={Oct}, pages={139–149} } @article{prince_chitwood_lashley_deperno_moorman_2016, title={Resource selection by southeastern fox squirrels in a fire-maintained forest system}, volume={97}, ISSN={["1545-1542"]}, DOI={10.1093/jmammal/gyv210}, abstractNote={Abstract}, number={2}, journal={JOURNAL OF MAMMALOGY}, author={Prince, Annemarie and Chitwood, M. Colter and Lashley, Marcus A. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2016}, month={Apr}, pages={631–638} } @article{stevenson_chitwood_lashley_pollock_swingen_moorman_deperno_2016, title={Survival and Cause-Specific Mortality of Coyotes on a Large Military Installation}, volume={15}, ISSN={["1938-5412"]}, DOI={10.1656/058.015.0307}, abstractNote={Abstract Canis latrans (Coyote) recently expanded into the southeastern United States, creating ecologically novel interactions with other species. However, relatively few studies have examined vital rates of southeastern Coyotes or estimated vital rates where individuals are protected from hunting and trapping. In 2011, we captured and attached GPS radiocollars to 31 Coyotes at Fort Bragg Military Installation, NC, where Coyote harvest was restricted. We used a 12-month period (February 2011–January 2012) and known-fate modeling in Program MARK to estimate annual survival. Model-selection results indicated the time-varying model (S[t]) was the most parsimonious model, and. annual survival was 0.80 (95% CI = 0.60–0.91). We documented 7 mortalities, including 2 from vehicles, 2 from offsite trapping, and 3 from unknown causes. Estimated Coyote survival rates at Fort Bragg were similar to most other estimates from the southeastern US. Anthropogenic causes of mortality were important even though hunting and trapping were restricted locally.}, number={3}, journal={SOUTHEASTERN NATURALIST}, author={Stevenson, Elizabeth R. and Chitwood, M. Colter and Lashley, Marcus A. and Pollock, Kenneth H. and Swingen, Morgan B. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2016}, month={Sep}, pages={459–466} } @article{grodsky_moorman_fritts_hazel_homyack_castleberry_wigley_2016, title={Winter bird use of harvest residues in clearcuts and the implications of forest bioenergy harvest in the southeastern United States}, volume={379}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2016.07.045}, abstractNote={Increased market viability of harvest residues gleaned for forest bioenergy feedstocks may intensify downed wood removal, particularly in intensively managed forests of the Southeast. Downed wood provides food and cover for many wildlife species, including birds, yet we are aware of no study that has examined winter bird response to experimentally manipulated, operational-scale woody biomass harvests. Further, little research has investigated avian use of downed wood following timber harvests. As such, our objectives were to: (1) evaluate effects of varying intensities of woody biomass harvest on the winter bird community and (2) document spatial associations between winter bird species and available habitat structure, including downed wood, in regenerating stands. In January and February of 2012–2014, we surveyed birds using a modified version of spot-mapping in six woody biomass removal treatments in North Carolina, USA (n = 4 regenerating stands). Treatments included clearcut harvest followed by: (1) traditional woody biomass harvest with no biomass harvesting guidelines; (2) 15% retention with harvest residues dispersed; (3) 15% retention with harvest residues clustered; (4) 30% retention with harvest residues dispersed; (5) 30% retention with harvest residues clustered; and (6) no woody biomass harvest (i.e., reference). We tested for treatment-level effects on avian relative abundance (overall and individual species), species diversity and richness, and counts of winter birds detected near (∼1 m from pile), in, or on branches of downed wood piles and calculated proportional avian habitat use of harvest residues and vegetation in regenerating stands. In 69 visits over three winters, we observed 3352 birds in treatments. In 2013, counts of birds detected in piles were greater in the no biomass harvest and 30% clustered treatments than the no biomass harvesting guidelines treatment. In 2012 and 2013 combined, mourning dove (Zenaida macroura) had greater relative abundance in the no biomass harvest treatment compared to the 15% dispersed treatment and was more often detected within 1 m of downed wood piles than in vegetation. We counted more winter birds in and near adjacent forest edge than in treatment interiors each year. Overall, we detected minimal treatment effects on winter bird relative abundance and no effects on species diversity and richness. Relative abundance of winter birds increased over time as vegetative cover established in regenerating stands. Our results suggest woody biomass harvests in intensively managed pine forests had no effect on the winter bird community, but winter birds used harvest residues. Further, vegetation structure and composition, rather than availability of harvest residues, primarily influenced winter bird use of regenerating stands.}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Grodsky, Steven M. and Moorman, Christopher E. and Fritts, Sarah R. and Hazel, Dennis W. and Homyack, Jessica A. and Castleberry, Steven B. and Wigley, T. Bently}, year={2016}, month={Nov}, pages={91–101} } @misc{marshall_orr_bradley_moorman_2015, title={A Review of Organic Lawn Care Practices and Policies in North America and the Implications of Lawn Plant Diversity and Insect Pest Management}, volume={25}, ISSN={["1943-7714"]}, DOI={10.21273/horttech.25.4.437}, abstractNote={There are ≈40 million acres of turfgrass lawns throughout the United States, most of which are managed under chemical-intensive pest and fertilizer programs. “Organic lawn care” is being adopted more widely; however, unlike the formally defined policies and regulations that govern organic agriculture, the label organic lawn management has not been formally defined and is used to describe a variety of practices. Neighborhoods, cities, states, and provinces across North America are adopting policies regulating the use of pesticides and fertilizers in the landscape. In addition, a small but growing number of public institutions and individual consumers are successfully adopting alternative lawn care methods, including organic lawn care. Although perceived as environmentally friendly, the effects of organic management on insect diversity and pest management remain understudied. Organic lawn management may lead to increased lawn plant diversity, which in agroecosystems has enhanced ecological services provided by beneficial insect species. Effects of vegetative diversity on lawn pest management are less clear. Vegetative complexity and increased plant diversity in urban landscapes may enhance insect predator efficacy. The diversity of predatory insects varies between turfgrass varieties in response to prey populations. Mortality of insectivorous and granivorous ground beetles (Carabidae) while not directly impacted by pest management programs in turfgrass may be indirectly impacted by a reduction in the prevalence of plant species that provide alternative food resources. Previous studies have focused on herbivorous insects as well as predatory and parasitic insects that feed on them. Future studies should assess how lawn plant diversity resulting from organic management practices might impact insect communities in turfgrass.}, number={4}, journal={HORTTECHNOLOGY}, author={Marshall, Sam and Orr, David and Bradley, Lucy and Moorman, Christopher}, year={2015}, month={Aug}, pages={437–446} } @article{serenari_peterson_moorman_cubbage_jervis_2015, title={Application of Choice Experiments to Determine Stakeholder Preferences for Woody Biomass Harvesting Guidelines}, volume={34}, ISSN={1054-9811 1540-756X}, url={http://dx.doi.org/10.1080/10549811.2015.1007511}, DOI={10.1080/10549811.2015.1007511}, abstractNote={Biomass harvesting guidelines (BHGs) have been developed to address concerns about the sustainability of harvesting woody biomass. Assessing preferences among BHG stakeholders is important for designing operationally feasible and socially acceptable standards in different contexts. We used choice modeling to determine how foresters, loggers, and landowners perceived the relative importance of stumpage price, wildlife habitat quality, percentage of coarse woody debris (CWD) remaining, and distribution of CWD in their choices of BHG scenarios. Responses (N = 718) indicated stumpage price was nearly double the importance of wildlife habitat quality, and three times more important than debris distribution and debris remaining.}, number={4}, journal={Journal of Sustainable Forestry}, publisher={Informa UK Limited}, author={Serenari, Christopher and Peterson, M. Nils and Moorman, Christopher E. and Cubbage, Frederick and Jervis, Suzanne}, year={2015}, month={Feb}, pages={343–357} } @article{chitwood_lashley_kilgo_pollock_moorman_deperno_2015, title={Do Biological and Bedsite Characteristics Influence Survival of Neonatal White-Tailed Deer?}, volume={10}, ISSN={1932-6203}, url={http://dx.doi.org/10.1371/journal.pone.0119070}, DOI={10.1371/journal.pone.0119070}, abstractNote={Coyotes recently expanded into the eastern U.S. and potentially have caused localized white-tailed deer population declines. Research has focused on quantifying coyote predation on neonates, but little research has addressed the potential influence of bedsite characteristics on survival. In 2011 and 2012, we radiocollared 65 neonates, monitored them intensively for 16 weeks, and assigned mortality causes. We used Program MARK to estimate survival to 16 weeks and included biological covariates (i.e., sex, sibling status [whether or not it had a sibling], birth weight, and Julian date of birth). Survival to 16 weeks was 0.141 (95% CI = 0.075-0.249) and the top model included only sibling status, which indicated survival was lower for neonates that had a sibling. Predation was the leading cause of mortality (35 of 55; 64%) and coyotes were responsible for the majority of depredations (30 of 35; 86%). Additionally, we relocated neonates for the first 10 days of life and measured distance to firebreak, visual obstruction, and plant diversity at bedsites. Survival of predation to 10 days (0.726; 95% CI = 0.586-0.833) was weakly associated with plant diversity at bedsites but not related to visual obstruction. Our results indicate that neonate survival was low and coyote predation was an important source of mortality, which corroborates several recent studies from the region. Additionally, we detected only weak support for bedsite cover as a covariate to neonate survival, which indicates that mitigating effects of coyote predation on neonates may be more complicated than simply managing for increased hiding cover.}, number={3}, journal={PLOS ONE}, publisher={Public Library of Science (PLoS)}, author={Chitwood, M. Colter and Lashley, Marcus A. and Kilgo, John C. and Pollock, Kenneth H. and Moorman, Christopher E. and DePerno, Christopher S.}, editor={Roca, Alfred L.Editor}, year={2015}, month={Mar}, pages={e0119070} } @article{raybuck_moorman_fritts_greenberg_deperno_simon_warburton_2015, title={Do silvicultural practices to restore oaks affect salamanders in the short term?}, volume={21}, ISSN={["1903-220X"]}, DOI={10.2981/wlb.00076}, abstractNote={Salamanders are an important ecological component of eastern hardwood forests and may be affected by natural or silvicultural disturbances that alter habitat structure and associated microclimate. From May to August in 2008 (pretreatment) and 2011 (post‐treatment), we evaluated the response of salamanders to three silvicultural practices designed to promote oak regeneration — prescribed fire, midstory herbicide application and shelterwood harvest — and a control. We trapped salamanders using drift fences with pitfall traps in five replicates of the four treatments. Only the southern gray‐cheeked salamander Plethodon metcalfi and the southern Appalachian salamander P. teyahalee were captured in sufficient numbers for robust statistical analysis. We analyzed data for these species using single‐species dynamic occupancy models in statistical software program R. We allowed changes in four covariates to influence extinction probability from pre‐ to post‐treatment implementation: 1) percent leaf litter cover; 2) percent understory cover; 3) percent CWD cover; and 4) percent canopy cover. The final combined model set describing extinction probability contained four models with ΔAIC < 2 for P. metcalfi and nine models with ΔAIC < 2, including the null model, for P. teyahalee. For both species, the 95% confidence intervals for model‐averaged extinction probability parameter estimates overlapped zero, suggesting none were significant predictors of extinction probability. Absence of short‐term salamander response in midstory herbicide and prescribed burn treatments was likely because of minor or transitory changes to forest structure. In shelterwood harvests, any potential effects of reduced canopy and leaf litter cover may have been mitigated by rapid post‐treatment vegetation sprouting. Additionally, climatic conditions associated with high elevation sites and high amounts of rainfall in 2011 may have compensated for potential changes to microclimate. Continued monitoring of Plethodon salamanders to assess responses at longer time scales (e.g. > 3 years post‐treatment) is warranted.}, number={4}, journal={WILDLIFE BIOLOGY}, author={Raybuck, Amy L. and Moorman, Christopher E. and Fritts, Sarah R. and Greenberg, Cathryn H. and Deperno, Christopher S. and Simon, Dean M. and Warburton, Gordon S.}, year={2015}, month={Aug}, pages={186–194} } @article{rutledge_moorman_washburn_deperno_2015, title={Evaluation of Resident Canada Goose Movements to Reduce the Risk of Goose-Aircraft Collisions at Suburban Airports}, volume={79}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.924}, abstractNote={ABSTRACT}, number={7}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Rutledge, M. Elizabeth and Moorman, Christopher E. and Washburn, Brian E. and Deperno, Christopher S.}, year={2015}, month={Sep}, pages={1185–1191} } @article{stevenson_lashley_chitwood_peterson_moorman_2015, title={How Emotion Trumps Logic in Climate Change Risk Perception: Exploring the Affective Heuristic Among Wildlife Science Students}, volume={20}, ISSN={1087-1209 1533-158X}, url={http://dx.doi.org/10.1080/10871209.2015.1077538}, DOI={10.1080/10871209.2015.1077538}, abstractNote={Despite scientific support for the reality of climate change, public opinion remains polarized. Continued skepticism may be partially explained by lack of understanding of climate change science, and worldview and ideology, but factors contributing to risk perceptions also may differ depending on the subject of risk. This article compared how wildlife students in the eastern United States perceive climate change risk to wildlife versus humans. Left-leaning political ideology and acceptance of anthropogenic global warming predicted perceptions of climate change risks to humans. Contrastingly, scientific understanding was the most important predictor of wildlife-related risk perceptions. Students may have used an affective heuristic (i.e., emotions) in assessing climate change risks to society and a cognitive reasoning (i.e., logic and data) when considering climate change risks to wildlife, which suggests that climate change communicators should appeal to these different modes of thinking when considering risks to humans versus wildlife.}, number={6}, journal={Human Dimensions of Wildlife}, publisher={Informa UK Limited}, author={Stevenson, Kathryn T. and Lashley, Marcus A. and Chitwood, M. Colter and Peterson, M. Nils and Moorman, Christopher E.}, year={2015}, month={Oct}, pages={501–513} } @article{chitwood_swingen_lashley_flowers_palamar_apperson_olfenbutte_moorman_depernol_2015, title={PARASITOLOGY AND SEROLOGY OF FREE-RANGING COYOTES (CANIS LATRANS) IN NORTH CAROLINA, USA}, volume={51}, ISSN={["1943-3700"]}, DOI={10.7589/2015-01-002}, abstractNote={Abstract Coyotes (Canis latrans) have expanded recently into the eastern US and can serve as a source of pathogens to domestic dogs (Canis lupus familiaris), livestock, and humans. We examined free-ranging coyotes from central North Carolina, US, for selected parasites and prevalence of antibodies against viral and bacterial agents. We detected ticks on most (81%) coyotes, with Amblyomma americanum detected on 83% of those with ticks. Fifteen (47%) coyotes were positive for heartworms (Dirofilaria immitis), with a greater detection rate in adults (75%) than juveniles (22%). Serology revealed antibodies against canine adenovirus (71%), canine coronavirus (32%), canine distemper virus (17%), canine parvovirus (96%), and Leptospira spp. (7%). We did not detect antibodies against Brucella abortus/suis or Brucella canis. Our results showed that coyotes harbor many common pathogens that present health risks to humans and domestic animals and suggest that continued monitoring of the coyote's role in pathogen transmission is warranted.}, number={3}, journal={JOURNAL OF WILDLIFE DISEASES}, author={Chitwood, M. Colter and Swingen, Morgan B. and Lashley, Marcus A. and Flowers, James R. and Palamar, Maria B. and Apperson, Charles S. and Olfenbutte, Colleen and Moorman, Christopher E. and DePernol, Christopher S.}, year={2015}, month={Jul}, pages={664–669} } @article{lashley_chitwood_harper_moorman_deperno_2015, title={Poor soils and density-mediated body weight in deer: forage quality or quantity?}, volume={21}, ISSN={["1903-220X"]}, DOI={10.2981/wlb.00073}, abstractNote={Cervid studies have demonstrated body weight and lactation may be limited in areas with poor‐quality soils, with the underlying mechanism often attributed to poor forage quality resulting from poor soil quality. However, if highly nutritious foods are produced but in limited quantities, selective foraging may alleviate nutritional stressors associated with poor soil productivity when adequate quantities of high‐quality forage are obtained. We tested whether poor soil productivity limits forage from being high quality or conversely limits the abundance of high‐quality forages. To do so, we determined whether nutrient concentrations in selected and non‐selected forages on our poor soil study site met the nutritional requirements of lactating white‐tailed deer Odocoileus virginianus assuming adequate amounts of forage are obtained. Also, we compared body weight of yearling males at a high density (13–17 deer km‐2) and low density (3–5 deer km‐2), because previous studies concluded soils on the study site were too poor for morphometrics (e.g. body weight) to be density‐mediated. Deer selected plant species from each of five forage classes that would meet their nutritional requirements (i.e. assuming adequate forage intake) but also selected for different nutrients across forage classes. Phosphorus was limited in most forages, but deer selected forages that met P‐requirements 10 times more than expected. We demonstrated body weight was 7.3% greater when deer density was low than high. Contrary to previous reports from poor productivity soil regions, deer on our study site should be able to meet lactation requirements when the quantity of high‐quality forage is not limiting, and similarly body weight should be density‐mediated. Hence, management strategies that alter the amount of forage per animal (i.e. decreasing animal density and/or increasing forage abundance) are viable options to promote desirable ungulate morphometrics in poor soil regions.}, number={4}, journal={WILDLIFE BIOLOGY}, author={Lashley, Marcus A. and Chitwood, M. Colter and Harper, Craig A. and Moorman, Chris E. and DePerno, Chris S.}, year={2015}, month={Aug}, pages={213–219} } @article{lashley_chitwood_kays_harper_deperno_moorman_2015, title={Prescribed fire affects female white-tailed deer habitat use during summer lactation}, volume={348}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/j.foreco.2015.03.041}, DOI={10.1016/j.foreco.2015.03.041}, abstractNote={Prescribed fire commonly is used to manage habitat for white-tailed deer (Odocoileus virginianus). Although the effects of fire on forage availability for deer have been studied, how female deer use burned areas is not well known, particularly as it relates to fire season and the years-since-fire. We used GPS tracking data from 16 adult female white-tailed deer to assess the effects of fire season and years-since-fire on habitat use during summer lactation. Females selected unburned drainages and older (>1 yr-since-fire) burned areas, and avoided recently burned areas. Individuals with a greater percentage of their summer core area burned expanded the size of their summer home range but did not change summer core area size. Furthermore, summer core area site fidelity (i.e., % overlap between 2011 and 2012 core areas) decreased as the percentage of the 2011 summer core area burned in 2012 increased. Female deer increased selection of burned areas as years-since-fire increased, likely because there was a temporary loss of cover immediately following fire with plants slowly regenerating the subsequent growing seasons. Likewise, to avoid areas depleted of cover, females shifted their core areas away from recent burns when possible but increased their core area size when burned areas were unavoidable (i.e., a large portion of their home range was burned). Burning large contiguous areas may initially have a negative effect on female deer during lactation because of the depletion of cover.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Lashley, Marcus A. and Chitwood, M. Colter and Kays, Roland and Harper, Craig A. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2015}, month={Jul}, pages={220–225} } @article{fritts_grodsky_hazel_homyack_castleberry_moorman_2015, title={Quantifying multi-scale habitat use of woody biomass by southern toads}, volume={346}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/J.FORECO.2015.03.004}, DOI={10.1016/j.foreco.2015.03.004}, abstractNote={Woody biomass extraction for use as a feedstock for renewable energy may remove woody debris that provides suitable micro-climates for amphibians. We examined habitat use of the southern toad (Anaxyrus terrestris) as an indicator of relationships between amphibians and woody biomass in pine plantations of the southeastern United States using a controlled enclosure experiment and a field-based radio-telemetry study. In the enclosure experiment, we recorded toad selection among four 16-m2 treatments that varied in area of ground surface covered by coarse woody debris (CWD) and spatial allocation of CWD. Treatments were: (1) ≈100% of the ground area covered by CWD in one large pile (volume of CWD = 1.10 m3, 100CWD); (2) ≈50% of the ground area covered with CWD in one large pile (volume of CWD = 0.60 m3, 50PILE); (3) ≈50% of the ground area covered with CWD dispersed throughout the treatment (volume of CWD = 0.25 m3, 50DISP); and (4) no CWD (0CWD). In the radio-telemetry study, we identified southern toad daytime refuge locations and compared habitat characteristics to paired random locations. From May to August 2013, toads (n = 47) did not use enclosure treatments randomly during nocturnal hours (P < 0.01), and ranking of treatments from most to least selected was 0CWD, 100CWD, 50DISP, 50PILE. When no rain events occurred, toads spent a greater proportion of time during nocturnal hours in 100CWD as temperature increased (P = 0.03). Toads used 100CWD 75% of the time for diurnal refuge. Radio-marked toads (n = 37) avoided grass (P < 0.01) and bare ground (P < 0.01) as diurnal refuge sites. Although radio-marked toads used CWD, other cover sources also were used as refuge sites and toads did not select CWD cover (P = 0.11) over other diurnal refuge types. Our results suggest woody biomass in recently harvested pine plantations is not an essential habitat characteristic during nocturnal hours and therefore may not be important for foraging. Yet, woody biomass may provide diurnal refuge for southern toads, and likely other amphibians, when desiccation risk is high (i.e., temperatures are high and rain does not occur). Additionally, southern toads may use woody biomass for diurnal refuge when other cover sources are not available, but can exhibit behavioral plasticity when cover sources such as vegetation are accessible.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Fritts, S.R. and Grodsky, S.M. and Hazel, D.W. and Homyack, J.A. and Castleberry, S.B. and Moorman, C.E.}, year={2015}, month={Jun}, pages={81–88} } @article{swingen_deperno_moorman_2015, title={Seasonal Coyote Diet Composition at a Low-Productivity Site}, volume={14}, ISSN={["1938-5412"]}, DOI={10.1656/058.014.0219}, abstractNote={Abstract Canis latrans (Coyote) recently expanded its range into the southeastern US, where local data on Coyote diets are lacking. We studied Coyote diets in a low-productivity area where food resources may be scarce. We determined Coyote diet composition through analysis of 315 scats collected at Fort Bragg Military Installation, Fort Bragg, NC, between May 2011 and July 2012. Odocoileus virginianus (White-tailed Deer) was the most common mammalian food item, occurring in 14.9% of all scats and 42.5% of winter scats. Soft-mast occurrence in Coyote diets was greatest in the fall, when Diospyros virginiana (Persimmon) occurred in most Coyote scats (95.7%). Coyotes on our low-productivity study site shifted their diets throughout the year based on the availability of food items and had a diet diversity similar to what has been reported for animals elsewhere in the species' range.}, number={2}, journal={SOUTHEASTERN NATURALIST}, author={Swingen, Morgan B. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2015}, month={Jun}, pages={397–404} } @article{grodsky_iglay_sorenson_moorman_2015, title={Should Invertebrates Receive Greater Inclusion in Wildlife Research Journals?}, volume={79}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.875}, abstractNote={ABSTRACT}, number={4}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Grodsky, Steven M. and Iglay, Raymond B. and Sorenson, Clyde E. and Moorman, Christopher E.}, year={2015}, month={May}, pages={529–536} } @article{fritts_moorman_grodsky_hazel_homyack_farrell_castleberry_2015, title={Shrew response to variable woody debris retention: Implications for sustainable forest bioenergy}, volume={336}, ISSN={0378-1127}, url={http://dx.doi.org/10.1016/J.FORECO.2014.10.009}, DOI={10.1016/j.foreco.2014.10.009}, abstractNote={Shrews are integral components of forest food webs and may rely on downed woody debris to provide microhabitats that satisfy high moisture and metabolic requirements. However, woody biomass harvests glean downed woody debris to use as a bioenergy feedstock. Biomass Harvesting Guidelines (BHGs) provide guidance on the amount and distribution of downed woody debris retained after harvest to ensure ecological sustainability of woody biomass harvesting and limit detrimental effects on wildlife. However, the success of Biomass Harvesting Guidelines at reaching sustainability goals, including conservation of wildlife habitat, has not been tested in an operational setting. Thus, we compared shrew captures among six woody biomass harvesting treatments in pine plantations in North Carolina, USA from April to August 2011–2014 (n = 4) and Georgia, USA from April to August 2011–2013 (n = 4). Treatments included: (1) woody biomass harvest with no BHGs; (2) 15% retention with woody biomass dispersed; (3) 15% retention with woody biomass clustered; (4) 30% retention with woody biomass dispersed; (5) 30% retention with woody biomass clustered; and (6) no woody biomass harvested. We sampled shrews with drift fence arrays and compared relative abundance of shrews among treatments using analysis of variance. Additionally, we used general linear regression models to evaluate the influence of downed woody debris volume and vegetation structure on shrew capture success at each drift fence for species with >100 captures/state/year. In 53,690 trap nights, we had 1,712 shrew captures representing three species, Cryptotis parva, Blarina carolinensis, and Sorex longirostris. We did not detect consistent differences in shrew relative abundance among woody biomass harvest treatments, but relative abundance of all species increased over time as vegetation became established. In North Carolina, total shrew capture success was negatively related to volume of downed woody debris within 50 m of the drift fence array (P = 0.05) in 2013 and positively related to bare groundcover in 2013 (P = 0.02) and 2014 (P < 0.01). In Georgia, total shrew capture success was negatively related to herbaceous groundcover (P < 0.01) and leaf litter groundcover (P = 0.02) and positively related to woody vegetation groundcover (P < 0.01) and vertical vegetation structure (P = 0.03) in 2013. Our results suggest that shrews in our study area were associated more with vegetation characteristics than downed woody debris and that woody biomass harvests may have little influence on shrew abundances in the southeastern United States Coastal Plain.}, journal={Forest Ecology and Management}, publisher={Elsevier BV}, author={Fritts, S.R. and Moorman, C.E. and Grodsky, S.M. and Hazel, D.W. and Homyack, J.A. and Farrell, C.B. and Castleberry, S.B.}, year={2015}, month={Jan}, pages={35–43} } @article{klimstra_moorman_converse_royle_harper_2015, title={Small Mammal Use of Native Warm-Season and Non-Native Cool-Season Grass Forage Fields}, volume={39}, ISSN={["1938-5463"]}, DOI={10.1002/wsb.507}, abstractNote={ABSTRACT}, number={1}, journal={WILDLIFE SOCIETY BULLETIN}, author={Klimstra, Ryan L. and Moorman, Christopher E. and Converse, Sarah J. and Royle, J. Andrew and Harper, Craig A.}, year={2015}, month={Mar}, pages={49–55} } @article{taillie_peterson_moorman_2015, title={The relative importance of multiscale factors in the distribution of Bachman's Sparrow and the implications for ecosystem conservation}, volume={117}, ISSN={["1938-5129"]}, DOI={10.1650/condor-14-137.1}, abstractNote={ABSTRACT Recent research has shown that landscape-level changes, namely habitat loss and fragmentation, can play an important role in determining the distribution of species across a variety of ecological systems. However, the influence of these large-scale factors in relation to small-scale factors, such as local vegetation structure or composition, is poorly understood. We used Bachman's Sparrow (Peucaea aestivalis) as a surrogate species to measure the relative importance of local vegetation and large-scale habitat distribution in the Onslow Bight region of North Carolina, USA. We conducted repeated point counts at 232 points within 111 habitat patches between April 10 and July 20, 2011. We then fit a series of single-season occupancy models, including both local and landscape-level predictors, to identify those that best explained the distribution of Bachman's Sparrows. We documented a strong response to vegetation characteristics best maintained via prescribed fire, but the most influential predictor of Bachman's Sparrow occupancy was the amount of habitat within 3 km. Specifically, the probability of Bachman's Sparrow occurrence was close to zero in landscapes comprised of <10% habitat, regardless of local vegetation conditions. Our results illustrate the strong influence of habitat loss on Bachman's Sparrow and likely on other members of this community, many of which are of high conservation concern.}, number={2}, journal={CONDOR}, author={Taillie, Paul J. and Peterson, M. Nils and Moorman, Christopher E.}, year={2015}, month={May}, pages={137–146} } @article{lashley_chitwood_harper_deperno_moorman_2015, title={VARIABILITY IN FIRE PRESCRIPTIONS TO PROMOTE WILDLIFE FOODS IN THE LONGLEAF PINE ECOSYSTEM}, volume={11}, ISSN={["1933-9747"]}, DOI={10.4996/fireecology.1103062}, abstractNote={Abstract}, number={3}, journal={FIRE ECOLOGY}, author={Lashley, Marcus A. and Chitwood, M. Colter and Harper, Craig A. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2015}, pages={62–79} } @article{chitwood_lashley_kilgo_moorman_deperno_2015, title={White-Tailed Deer Population Dynamics and Adult Female Survival in the Presence of a Novel Predator}, volume={79}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.835}, abstractNote={ABSTRACT}, number={2}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Chitwood, Michael C. and Lashley, Marcus A. and Kilgo, John C. and Moorman, Christopher E. and Deperno, Christopher S.}, year={2015}, month={Feb}, pages={211–219} } @article{fritts_moorman_hazel_jackson_2014, title={Biomass Harvesting Guidelines affect downed woody debris retention}, volume={70}, ISSN={["1873-2909"]}, DOI={10.1016/j.biombioe.2014.08.010}, abstractNote={Our objective was to determine if a retention area-based Biomass Harvesting Guideline (BHG) strategy maintained desired volumes of downed woody debris (DWD) following woody biomass harvests. We implemented six randomly-assigned treatments in four clearcuts in loblolly pine plantations in the Coastal Plain physiographic region of North Carolina during 2010–2011: 1) woody biomass harvest with no BHGs (NOBHG); 2) 15% retention with woody biomass dispersed (15DISP); 3) 15% retention with woody debris clustered (15CLUS); 4) 30% retention with woody biomass dispersed (30DISP); 5) 30% retention with woody biomass clustered (30CLUS); and 6) no woody biomass harvest (i.e., clearcut only; NOBIOHARV). Prior to harvesting, we flagged 15% or 30% of the treatment area to serve as woody biomass retention sources for the four BHG treatments, and all woody biomass from the flagged area were retained and distributed across that entire treatment area. We examined effects of treatments on: 1) fraction estimated volume of pre-harvest standing volume (total and woody biomass) retained as residual DWD; and 2) fraction retained DWD in treatments 2–5 relative to retained DWD in the NOBHGS and NOBIOHARV treatments. Adding a woody biomass harvest reduced volume of residual DWD by 81% in NOBHG compared to NOBIOHARV. Estimates based on the second metric were most similar to target retentions with retention percentages at 18.8% in 15CLUS, 14.1% in 15DISP, 39.0% in 30CLUS, and 38.0% in 30DISP. Treatments resulted in retention of DWD fractions approximate to those prescribed, suggesting BHGs can be implemented successfully in an operational setting.}, journal={BIOMASS & BIOENERGY}, author={Fritts, S. R. and Moorman, C. E. and Hazel, D. W. and Jackson, B. D.}, year={2014}, month={Nov}, pages={382–391} } @article{chitwood_lashley_moorman_deperno_2014, title={Confirmation of Coyote Predation on Adult Female White-tailed Deer in the Southeastern United States}, volume={13}, ISSN={["1938-5412"]}, DOI={10.1656/058.013.0316}, abstractNote={Abstract The recent expansion of Canis latrans (Coyote) into the eastern United States has generated interest among wildlife managers because of the potential impacts on Odocoileus virginianus (White-tailed Deer) populations. Coyotes have been reported as predators of adult and neonate White-tailed Deer in some parts of their range, but recent studies in the Southeast have documented only Coyote predation on neonates. We report 4 confirmed Coyote predation events on adult female White-tailed Deer that were radiocollared, implanted with vaginal implant transmitters, monitored every 4–8 hours, and apparently healthy. Field necropsies confirmed killing-bite wounds to the upper throat and base of the mandible, and feeding behavior on the carcasses was consistent with what has been observed for Coyotes. Further, we used swabs from bite wounds to confirm the presence of predator DNA, and the 3 carcasses that were swabbed tested positive for the presence of Coyote DNA. To our knowledge, our results represent the first scientifically documented Coyote predations on adult female White-tailed Deer in the Southeast.}, number={3}, journal={SOUTHEASTERN NATURALIST}, author={Chitwood, M. Colter and Lashley, Marcus A. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2014}, month={Sep}, pages={N30–N32} } @article{charles-smith_rutledge_meek_baine_massey_ellsaesser_deperno_moorman_degernes_2014, title={Hematologic Parameters and Hemoparasites of Nonmigratory Canada Geese (Branta canadensis) From Greensboro, North Carolina, USA}, volume={28}, ISSN={["1938-2871"]}, DOI={10.1647/2012-072}, abstractNote={Abstract Large flocks of wild, nonmigratory Canada geese (Branta canadensis) have established permanent residence throughout the eastern United States and have become a public concern. Few studies have assessed the hematologic parameters for these populations, which could provide useful information for monitoring individual and population health of Canada geese. This study measured the hematologic parameters and detected the presence of hemoparasites from 146 wild, nonmigratory Canada geese in central North Carolina, USA, during their annual molt. The age class, sex, and weight of each bird were recorded at capture. Values for packed cell volume (PCV), estimated white blood cell count, white blood cell differentials, and heterophil : lymphocyte ratios were calculated for each bird. Adults and female geese had higher estimated white blood cell counts compared with juveniles and males, respectively. The PCV increased with weight and age class. Adult geese had higher percentages of heterophils and heterophil : lymphocyte ratios, whereas juvenile geese had higher percentages of lymphocytes. Relative eosinophil counts in adults increased with decreasing bird weight, and relative monocyte counts in juveniles increased with increasing weight. Three percent of geese were infected with species of Hemoproteus blood parasites. Atypical lymphocyte morphology, including pseudopods, split nuclei, and cytoplasmic granules, was observed in 5% of the birds. The hematologic values reported for adult and juvenile nonmigratory Canada geese in this study may serve as reference intervals for ecological studies and veterinary care of wild and captive Canada geese.}, number={1}, journal={JOURNAL OF AVIAN MEDICINE AND SURGERY}, author={Charles-Smith, Lauren E. and Rutledge, M. Elizabeth and Meek, Caroline J. and Baine, Katherine and Massey, Elizabeth and Ellsaesser, Laura N. and DePerno, Christopher S. and Moorman, Christopher E. and Degernes, Laurel A.}, year={2014}, month={Mar}, pages={16–23} } @article{bowling_moorman_deperno_gardner_2014, title={Influence of Landscape Composition on Northern Bobwhite Population Response to Field Border Establishment}, volume={78}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.639}, abstractNote={ABSTRACT}, number={1}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Bowling, Shannon A. and Moorman, Christopher E. and Deperno, Christopher S. and Gardner, Beth}, year={2014}, month={Jan}, pages={93–100} } @article{garabedian_mcgaughey_reutebuch_parresol_kilgo_moorman_peterson_2014, title={Quantitative analysis of woodpecker habitat using high-resolution airborne LiDAR estimates of forest structure and composition}, volume={145}, ISSN={0034-4257}, url={http://dx.doi.org/10.1016/J.RSE.2014.01.022}, DOI={10.1016/j.rse.2014.01.022}, abstractNote={Light detection and ranging (LiDAR) technology has the potential to radically alter the way researchers and managers collect data on wildlife–habitat relationships. To date, the technology has fostered several novel approaches to characterizing avian habitat, but has been limited by the lack of detailed LiDAR-habitat attributes relevant to species across a continuum of spatial grain sizes and habitat requirements. We demonstrate a novel three-step approach for using LiDAR data to evaluate habitat based on multiple habitat attributes and accounting for their influence at multiple grain sizes using federally endangered red-cockaded woodpecker (RCW; Picoides borealis) foraging habitat data from the Savannah River Site (SRS) in South Carolina, USA. First, we used high density LiDAR data (10 returns/m2) to predict detailed forest attributes at 20-m resolution across the entire SRS using a complementary application of nonlinear seemingly unrelated regression and multiple linear regression models. Next, we expanded on previous applications of LiDAR by constructing 95% joint prediction confidence intervals to quantify prediction error at various spatial aggregations and habitat thresholds to determine a biologically and statistically meaningful grain size. Finally, we used aggregations of 20-m cells and associated confidence interval boundaries to demonstrate a new approach to produce maps of RCW foraging habitat conditions based on the guidelines described in the species' recovery plan. Predictive power (R2) of regression models developed to populate raster layers ranged from 0.34 to 0.81, and prediction error decreased as aggregate size increased, but minimal reductions in prediction error were observed beyond 0.64-ha (4 × 4 20-m cells) aggregates. Mapping habitat quality while accounting for prediction error provided a robust method to determine the potential range of habitat conditions and specific attributes that were limiting in terms of the amount of suitable habitat. The sequential steps of our analytical approach provide a useful framework to extract detailed and reliable habitat attributes for a forest-dwelling habitat specialist, broadening the potential to apply LiDAR in conservation and management of wildlife populations.}, journal={Remote Sensing of Environment}, publisher={Elsevier BV}, author={Garabedian, James E. and McGaughey, Robert J. and Reutebuch, Stephen E. and Parresol, Bernard R. and Kilgo, John C. and Moorman, Christopher E. and Peterson, M. Nils}, year={2014}, month={Apr}, pages={68–80} } @article{lashley_chitwood_prince_elfelt_kilburg_deperno_moorman_2014, title={Subtle effects of a managed fire regime: A case study in the longleaf pine ecosystem}, volume={38}, ISSN={["1872-7034"]}, DOI={10.1016/j.ecolind.2013.11.006}, abstractNote={Land managers often use fire prescriptions to mimic intensity, season, completeness, and return interval of historical fire regimes. However, fire prescriptions based on average historical fire regimes do not consider natural stochastic variability in fire season and frequency. Applying prescribed fire based on averages could alter the relative abundance of important plant species and structure. We evaluated the density and distribution of oak (Quercus spp.) and persimmon (Diospyros virgiana) stems and mast after 22 yr of a historical-based growing-season fire prescription that failed to consider the variability in historical fire regimes. We randomly established 30 25-m transects in each of 5 vegetation types and counted reproductively mature oak and persimmon stems and their fruits. In upland longleaf pine (Pinus palustris) stands, this fire regime killed young hardwood trees, thereby decreasing compositional and structural heterogeneity within the upland pine vegetation type and limiting occurrence of the upland hardwood vegetation type. Acorns and persimmons were disproportionately distributed near firebreaks within low intensity fire transition zones. Mast was maintained, though in an unnatural distribution, as a result of an elaborate firebreak system. Our data indicate managed fire regimes may fail to mimic spatial distribution, frequency, and intensity of historical disturbances even when the fire prescription is based on empirical reference fire regimes. To maximize structural heterogeneity and conserve key ecosystem functionality, fire prescriptions should include variations in frequency, season, application method, and fire weather conditions rather than focusing on an average historical fire regime.}, journal={ECOLOGICAL INDICATORS}, author={Lashley, Marcus A. and Chitwood, M. Colter and Prince, Annemarie and Elfelt, Morgan B. and Kilburg, Eric L. and DePerno, Christopher S. and Moorman, Christopher E.}, year={2014}, month={Mar}, pages={212–217} } @article{prince_deperno_gardner_moorman_2014, title={Survival and home-range size of southeastern fox squirrels in North Carolina}, volume={13}, number={3}, journal={Southeastern Naturalist}, author={Prince, A. and DePerno, C. S. and Gardner, B. and Moorman, C. E.}, year={2014}, pages={456–462} } @misc{garabedian_moorman_peterson_kilgo_2014, title={Systematic review of the influence of foraging habitat on red-cockaded woodpecker reproductive success}, volume={20}, ISSN={["1903-220X"]}, DOI={10.2981/wlb.13004}, abstractNote={Relationships between foraging habitat and reproductive success provide compelling evidence of the contribution of specific vegetative features to foraging habitat quality, a potentially limiting factor for many animal populations. For example, foraging habitat quality likely will gain importance in the recovery of the threatened red‐cockaded woodpecker Picoides borealis (RCW) in the USA as immediate nesting constraints are mitigated. Several researchers have characterized resource selection by foraging RCWs, but emerging research linking reproductive success (e.g. clutch size, nestling and fledgling production, and group size) and foraging habitat features has yet to be synthesized. Therefore, we reviewed peer‐refereed scientific literature and technical resources (e.g. books, symposia proceedings, and technical reports) that examined RCW foraging ecology, foraging habitat, or demography to evaluate evidence for effects of the key foraging habitat features described in the species' recovery plan on group reproductive success. Fitness‐based habitat models suggest foraging habitat with low to intermediate pine Pinus spp. densities, presence of large and old pines, minimal midstory development, and herbaceous groundcover support more productive RCW groups. However, the relationships between some foraging habitat features and RCW reproductive success are not well supported by empirical data. In addition, few regression models account for > 30% of variation in reproductive success, and unstandardized multiple and simple linear regression coefficient estimates typically range from ‐0.100 to 0.100, suggesting ancillary variables and perhaps indirect mechanisms influence reproductive success. These findings suggest additional research is needed to address uncertainty in relationships between foraging habitat features and RCW reproductive success and in the mechanisms underlying those relationships.}, number={1}, journal={WILDLIFE BIOLOGY}, author={Garabedian, James E. and Moorman, Christopher E. and Peterson, M. Nils and Kilgo, John C.}, year={2014}, month={Mar}, pages={37–46} } @article{rutledge_sollmann_washburn_moorman_deperno_2014, title={Using novel spatial mark-resight techniques to monitor resident Canada geese in a suburban environment}, volume={41}, ISSN={["1448-5494"]}, DOI={10.1071/wr14069}, abstractNote={Context Over the past two decades, an increase in the number of resident (non-migratory) Canada geese (Branta canadensis) in the United States has heightened the awareness of human–goose interactions. Aims Accordingly, baseline demographic estimates for goose populations are needed to help better understand the ecology of Canada geese in suburban areas. Methods As a basis for monitoring efforts, we estimated densities of adult resident Canada geese in a suburban environment by using a novel spatial mark–resight method. We resighted 763 neck- and leg-banded resident Canada geese two to three times per week in and around Greensboro, North Carolina, over an 18-month period (June 2008 – December 2009). We estimated the density, detection probabilities, proportion of male geese in the population, and the movements and home-range radii of the geese by season ((post-molt I 2008 (16 July – 31 October), post-molt II 2008/2009 (1 November – 31 January), breeding and nesting 2009 (1 February – 31 May), and post-molt I 2009). Additionally, we used estimates of the number of marked individuals to quantify apparent monthly survival. Key results Goose densities varied by season, ranging from 11.10 individuals per km2 (s.e. = 0.23) in breeding/nesting to 16.02 individuals per km2 (s.e. = 0.34) in post-molt II. The 95% bivariate normal home-range radii ranged from 2.60 to 3.86 km for males and from 1.90 to 3.15 km for females and female home ranges were smaller than those of male geese during the breeding/nesting and post-molt II seasons. Apparent monthly survival across the study was high, ranging from 0.972 (s.e. = 0.005) to 0.995 (s.e. = 0.002). Conclusions By using spatial mark–resight models, we determined that Canada goose density estimates varied seasonally. Nevertheless, the seasonal changes in density are reflective of the seasonal changes in behaviour and physiological requirements of geese. Implications Although defining the state–space of spatial mark–resight models requires careful consideration, the technique represents a promising new tool to estimate and monitor the density of free-ranging wildlife. Spatial mark–resight methods provide managers with statistically robust population estimates and allow insight into animal space use without the need to employ more costly methods (e.g. telemetry). Also, when repeated across seasons or other biologically important time periods, spatial mark–resight modelling techniques allow for inference about apparent survival.}, number={5}, journal={WILDLIFE RESEARCH}, author={Rutledge, M. Elizabeth and Sollmann, Rahel and Washburn, Brian E. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2014}, pages={447–453} } @article{chitwood_lashley_moorman_deperno_2014, title={Vocalization Observed in Starving White-tailed Deer Neonates}, volume={13}, ISSN={["1938-5412"]}, DOI={10.1656/058.013.0202}, abstractNote={Abstract We observed loud, frequent vocalizations by 5 Odocoileus virginianus (White-tailed Deer) neonates that ultimately died of starvation due to abandonment. We did not observe this behavior by other neonates, regardless of survival or cause of mortality. Thus, we believe that neonate vocalization could serve as a useful field indicator of abandonment. Additionally, estimates of predation rates may be inflated because they are masking high rates of undetected abandonment.}, number={2}, journal={SOUTHEASTERN NATURALIST}, author={Chitwood, M. Colter and Lashley, Marcus A. and Moorman, Christopher E. and DePerno, Christopher S.}, year={2014}, month={Jun}, pages={N6–N8} } @article{lashley_chitwood_biggerstaff_morina_moorman_deperno_2014, title={White-Tailed Deer Vigilance: The Influence of Social and Environmental Factors}, volume={9}, ISSN={1932-6203}, url={http://dx.doi.org/10.1371/journal.pone.0090652}, DOI={10.1371/journal.pone.0090652}, abstractNote={Vigilance behavior may directly affect fitness of prey animals, and understanding factors influencing vigilance may provide important insight into predator-prey interactions. We used 40,540 pictures taken withcamera traps in August 2011 and 2012to evaluate factors influencing individual vigilance behavior of white-tailed deer (Odocoileus virginianus) while foraging at baited sites. We used binary logistic regression to determine if individual vigilance was affected by age, sex, and group size. Additionally, we evaluated whether the time of the day,moon phase,and presence of other non-predatorwildlife species impacted individual vigilance. Juveniles were 11% less vigilant at baited sites than adults. Females were 46% more vigilant when fawns were present. Males and females spent more time feeding as group size increased, but with each addition of 1 individual to a group, males increased feeding time by nearly double that of females. Individual vigilance fluctuated with time of day andwith moon phase but generally was least during diurnal and moonlit nocturnal hours, indicating deer have the ability to adjust vigilance behavior to changing predation risk associated with varyinglight intensity.White-tailed deer increased individual vigilance when other non-predator wildlife were present. Our data indicate that differential effects of environmental and social constraints on vigilance behavior between sexes may encourage sexual segregation in white-tailed deer.}, number={3}, journal={PLoS ONE}, publisher={Public Library of Science (PLoS)}, author={Lashley, Marcus A and Chitwood, M. Colter and Biggerstaff, Michael T. and Morina, Daniel L. and Moorman, Christopher E. and DePerno, Christopher S.}, editor={Moreira, NeiEditor}, year={2014}, month={Mar}, pages={e90652} } @article{kilburg_moorman_deperno_cobb_harper_2014, title={Wild Turkey Nest Survival and Nest-Site Selection in the Presence of Growing-Season Prescribed Fire}, volume={78}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.751}, abstractNote={Concerns about destruction of wild turkey (Meleagris gallopavo) nests traditionally restricted the application of prescribed-fire to the dormant season in the southeastern United States. Periodic dormant- season burns were used to open forest understories and increase forage and nesting cover for wild turkeys. However, much of the Southeast historically burned during late spring and early summer (i.e., growing season), which tended to decrease understory woody vegetation and promote grasses and forbs, an important spring and summer food for wild turkeys. Despite the potential benefits of growing-season burns, landscape- scale application coincident with turkey nesting may destroy nests and reduce or redistribute woody nesting cover. We determined turkey nest-site selection and nest survival in a landscape managed with frequent growing-season burns. We monitored radio-tagged female wild turkeys to locate nests and determine nest survival. We compared vegetation composition and structure at nest sites to random sites within dominant cover types and calculated the probability of nest destruction as the product of the proportion of wild turkey nests active and the proportion of the landscape burned. Females selected shrub-dominated lowland ecotones (a transitional vegetation community between upland pine and bottomland hardwoods) for nesting and avoided upland pine. Ecotones had greater cover than upland pine and estimated nest survival in lowlands (60%) was greater than in uplands (10%). Although approximately 20% of the study area was burned concurrent with nesting activity, only 3.3% of monitored nests were destroyed by fire, and we calculated that no more than 6% of all turkey nests were exposed to fire annually on our study site. We suggest that growing- season burns have a minimal direct effect on turkey nest survival but may reduce nesting cover and structural and compositional heterogeneity in uplands, especially on poor quality soils. A combination of dormant and growing-season burns may increase nesting cover in uplands, while maintaining open stand conditions. 2014 The Wildlife Society.}, number={6}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Kilburg, Eric L. and Moorman, Christopher E. and Deperno, Christopher S. and Cobb, David and Harper, Craig A.}, year={2014}, month={Aug}, pages={1033–1039} } @article{moorman_plush_orr_reberg-horton_2013, title={Beneficial Insect Borders Provide Northern Bobwhite Brood Habitat}, volume={8}, ISSN={["1932-6203"]}, DOI={10.1371/journal.pone.0083815}, abstractNote={Strips of fallow vegetation along cropland borders are an effective strategy for providing brood habitat for declining populations of upland game birds (Order: Galliformes), including northern bobwhite (Colinus virginianus), but fallow borders lack nectar-producing vegetation needed to sustain many beneficial insect populations (e.g., crop pest predators, parasitoids, and pollinator species). Planted borders that contain mixes of prairie flowers and grasses are designed to harbor more diverse arthropod communities, but the relative value of these borders as brood habitat is unknown. We used groups of six human-imprinted northern bobwhite chicks as a bioassay for comparing four different border treatments (planted native grass and prairie flowers, planted prairie flowers only, fallow vegetation, or mowed vegetation) as northern bobwhite brood habitat from June-August 2009 and 2010. All field border treatments were established around nine organic crop fields. Groups of chicks were led through borders for 30-min foraging trials and immediately euthanized, and eaten arthropods in crops and gizzards were measured to calculate a foraging rate for each border treatment. We estimated arthropod prey availability within each border treatment using a modified blower-vac to sample arthropods at the vegetation strata where chicks foraged. Foraging rate did not differ among border treatments in 2009 or 2010. Total arthropod prey densities calculated from blower-vac samples did not differ among border treatments in 2009 or 2010. Our results showed plant communities established to attract beneficial insects should maximize the biodiversity potential of field border establishment by providing habitat for beneficial insects and young upland game birds.}, number={12}, journal={PLOS ONE}, publisher={Public Library of Science (PLoS)}, author={Moorman, Christopher E. and Plush, Charles J. and Orr, David B. and Reberg-Horton, Chris}, editor={Boyce, Mark S.Editor}, year={2013}, month={Dec} } @article{rutledge_siletzky_gu_degernes_moorman_deperno_kathariou_2013, title={CHARACTERIZATION OF CAMPYLOBACTER FROM RESIDENT CANADA GEESE IN AN URBAN ENVIRONMENT}, volume={49}, ISSN={["0090-3558"]}, DOI={10.7589/2011-10-287}, abstractNote={Waterfowl are natural reservoirs for zoonotic pathogens, and abundant resident (nonmigratory) Canada Geese (Branta canadensis) in urban and suburban environments pose the potential for transmission of Campylobacter through human contact with fecal deposits and contaminated water. In June 2008 and July 2009, we collected 318 fecal samples from resident Canada Geese at 21 locations in and around Greensboro, North Carolina, to test for Campylobacter. All campylobacter species detected were C. jejuni isolates, and prevalences in 2008 and 2009 were 5.0% and 16.0%, respectively. Prevalence of C. jejuni–positive sampling sites was 21% (3/14) and 40% (6/15) in 2008 and 2009, respectively. All C. jejuni isolates were susceptible to a panel of six antimicrobial agents (tetracycline, streptomycin, erythromycin, kanamycin, nalidixic acid, and ciprofloxacin). We used pulsed-field gel electrophoresis and fla-typing to identify several strain types among these isolates. Multilocus sequence typing of representative isolates revealed six sequence types, of which two (ST-3708 and ST-4368) were new, two (ST-702 and ST-4080) had been detected previously among C. jejuni from geese, and two (ST-991 and ST-4071) were first reported in C. jejuni from an environmental water source and a human illness, respectively. These results indicate a diverse population of antibiotic-susceptible C. jejuni in resident Canada Geese in and around Greensboro, North Carolina, and suggest a need for additional assessment of the public health risk associated with resident Canada Geese in urban and suburban areas.}, number={1}, journal={JOURNAL OF WILDLIFE DISEASES}, author={Rutledge, M. Elizabeth and Siletzky, Robin M. and Gu, Weimin and Degernes, Laurel A. and Moorman, Christopher E. and DePerno, Christopher S. and Kathariou, Sophia}, year={2013}, month={Jan}, pages={1–9} } @article{fox_reberg-horton_orr_moorman_frank_2013, title={Crop and field border effects on weed seed predation in the southeastern U.S. coastal plain}, volume={177}, ISSN={0167-8809}, url={http://dx.doi.org/10.1016/J.AGEE.2013.06.006}, DOI={10.1016/j.agee.2013.06.006}, abstractNote={Weed seed predation was studied in nine organic crop fields (three each of maize, soybeans and hay; 2.5–4.0 ha each) surrounded by four experimental field border types (planted native grass and prairie flowers, planted prairie flowers only, fallow vegetation, or mowed vegetation) during the fall of 2009 and 2010 in eastern North Carolina. We used predator exclusion cages to determine the amount of weed seed removal caused by invertebrates and vertebrates. Three common agricultural weed species, redroot pigweed (Amaranthus retroflexus), broadleaf signalgrass (Urochloa platyphylla), and sicklepod (Senna obtusifolia), were adhered to individual cards and placed inside the exclosure cages once a month for two weeks. Activity-density of invertebrate weed seed predators was measured with pitfall traps. Results show that field border type had no effect on seed removal rates, but that crop type heavily influenced both weed seed predation and invertebrate seed predator activity-density. Weed seed predation was highest in the dense, perennial hay fields and lowest in the more open harvested maize fields. Activity-densities for field crickets (Gryllus sp.) and the ground beetle Harpalus pennsylvanicus were also high in the hay fields and low in the maize fields, while the red imported fire ant (Solenopsis invicta) seemed to prefer the open maize fields. These results show that increasing vegetative diversity in field borders is not always an effective method for conserving weed seed predators, but that higher quality habitat inside the crop field can be achieved by increasing ground cover.}, journal={Agriculture, Ecosystems & Environment}, publisher={Elsevier BV}, author={Fox, Aaron F. and Reberg-Horton, S. Chris and Orr, David B. and Moorman, Christopher E. and Frank, Steven D.}, year={2013}, month={Sep}, pages={58–62} } @article{mcvey_cobb_powell_stoskopf_bobling_waits_moorman_2013, title={Diets of sympatric red wolves and coyotes in northeastern North Carolina}, volume={94}, ISSN={["1545-1542"]}, DOI={10.1644/13-mamm-a-109.1}, abstractNote={Abstract The recent co-occurrence of red wolves (Canis rufus) and coyotes (Canis latrans) in eastern North Carolina provides a unique opportunity to study prey partitioning by sympatric canids. We collected scats from this region and examined them for prey contents. We used fecal DNA analysis to identify which taxa deposited each scat and multinomial modeling designed for mark–recapture data to investigate diets of sympatric red wolves and coyotes. Diets of red wolves and coyotes did not differ, but the proportion of small rodents in the composite scats of both canids was greater in the spring than in the summer. White-tailed deer (Odocoileus virginianus), rabbits (Sylvilagus spp.), and small rodents were the most common diet items in canid scats. The similarity of diet between red wolves and coyotes suggests that these 2 species may be affecting prey populations similarly.}, number={5}, journal={JOURNAL OF MAMMALOGY}, author={McVey, Justin M. and Cobb, David T. and Powell, Roger A. and Stoskopf, Michael K. and Bobling, Justin H. and Waits, Lisette P. and Moorman, Christopher E.}, year={2013}, month={Oct}, pages={1141–1148} } @article{plush_moorman_orr_reberg-horton_2013, title={Overwintering sparrow use of field borders planted as beneficial insect habitat}, volume={77}, ISSN={0022-541X}, url={http://dx.doi.org/10.1002/jwmg.436}, DOI={10.1002/jwmg.436}, abstractNote={Abstract}, number={1}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Plush, Charles J. and Moorman, Christopher E. and Orr, David B. and Reberg-Horton, Chris}, year={2013}, month={Jan}, pages={200–206} } @article{allen_moorman_peterson_hess_moore_2013, title={Predicting success incorporating conservation subdivisions into land use planning}, volume={33}, ISSN={0264-8377}, url={http://dx.doi.org/10.1016/j.landusepol.2012.12.001}, DOI={10.1016/j.landusepol.2012.12.001}, abstractNote={Conservation subdivisions have emerged as a development option for communities wishing to conserve important ecological features and maintain rural character without decreasing housing density. Yet, these alternatives to conventional subdivisions rarely are used. We used logistic regression models to identify variables that predict county level success at adopting an ordinance and having a conservation subdivision built. Important predictors for adopting ordinances were median income, percent urban population, and a negative interaction between the two variables; important predictors for successfully completing a conservation subdivision were the adoption of an ordinance allowing conservation subdivisions and percent of residents with at least a four year college degree. Urban counties and the rural counties with higher median income were most successful adopting ordinances. Urban counties with higher education levels and an ordinance in place were most likely to have a conservation subdivision built within them. In poor rural counties, implementation may be more difficult because of limited resources to develop ordinances; these counties could collaborate with land trusts, other planning departments, or a regional council of governments to help lessen the financial burden associated with rewriting ordinances and implementing new land use practices.}, journal={Land Use Policy}, publisher={Elsevier BV}, author={Allen, Stephen and Moorman, Christopher and Peterson, M. Nils and Hess, George and Moore, Susan}, year={2013}, month={Jul}, pages={31–35} } @article{becker_moorman_deperno_simons_2013, title={Quantifiable Long-term Monitoring on Parks and Nature Preserves}, volume={12}, ISSN={["1938-5412"]}, DOI={10.1656/058.012.0208}, abstractNote={Abstract Herpetofauna have declined globally, and monitoring is a useful approach to document local and long-term changes. However, monitoring efforts often fail to account for detectability or follow standardized protocols. We performed a case study at Hemlock Bluffs Nature Preserve in Cary, NC to model occupancy of focal species and demonstrate a replicable long-term protocol useful to parks and nature preserves. From March 2010 to 2011, we documented occupancy of Ambystoma opacum (Marbled Salamander), Plethodon cinereus (Red-backed Salamander), Carphophis amoenus (Eastern Worm Snake), and Diadophis punctatus (Ringneck Snake) at coverboard sites and estimated breeding female Ambystoma maculatum (Spotted Salamander) abundance via dependent double-observer egg-mass counts in ephemeral pools. Temperature influenced detection of both Marbled and Red-backed Salamanders. Based on egg-mass data, we estimated Spotted Salamander abundance to be between 21 and 44 breeding females. We detected 43 of 53 previously documented herpetofauna species. Our approach demonstrates a monitoring protocol that accounts for factors that influence species detection and is replicable by parks or nature preserves with limited resources.}, number={2}, journal={SOUTHEASTERN NATURALIST}, author={Becker, Sharon and Moorman, Christopher and DePerno, Christopher and Simons, Theodore}, year={2013}, month={Jun}, pages={339–352} } @article{blackman_deperno_moorman_peterson_2013, title={Use of Crop Fields and Forest by Wintering American Woodcock}, volume={12}, ISSN={["1528-7092"]}, DOI={10.1656/058.012.0107}, abstractNote={Abstract - During the 1970s–80s, Scolopax minor (American Woodcock) on wintering grounds in North Carolina generally used bottomland forests diurnally and fed on earthworms in conventionally tilled soybean fields at night. Researchers surmised the ridges and furrows in conventionally tilled fields provided Woodcock protection from predators and winter weather. Since the 1980s, farmers widely adopted no-till practices for soybean agriculture, and this change in field structure may have altered Woodcock crop field use. We returned to the same area as previous research and conducted a study of Woodcock crop field and forest use in a landscape where crop fields are the dominant open-habitat type. During December 2009–March 2010, we captured and radio-tracked 29 Woodcock. Every 24 hours, we located each radio-marked Woodcock during diurnal and nocturnal periods, and verified the habitat type on foot as either crop field or bottomland forest. We recorded 94% of nocturnal locations in forest, 6% of nocturnal locations in crop fields, and 100% of diurnal locations in forest. Percent of an individual Woodcock's nocturnal locations in crop fields ranged from zero to 44%, with a mean of 6% (± 2% SE). The adoption of no-till technology and associated reduction in ridge and furrow micro-habitat available in crop fields may contribute to the low frequency of Woodcock nocturnal field use. Because Woodcock primarily were relocated in bottomland forests diurnally and nocturnally, forest stands should be conserved when managing agricultural landscapes.}, number={1}, journal={SOUTHEASTERN NATURALIST}, author={Blackman, Emily B. and DePerno, Christopher S. and Moorman, Christopher E. and Peterson, M. Nils}, year={2013}, month={Apr}, pages={85–92} } @article{moorman_bowen_kilgo_hanula_horn_ulyshen_2012, title={ARTHROPOD ABUNDANCE AND SEASONAL BIRD USE OF BOTTOMLAND FOREST HARVEST GAPS}, volume={124}, ISSN={["1938-5447"]}, DOI={10.1676/11-020.1}, abstractNote={Abstract We investigated the influence of arthropod abundance and vegetation structure on shifts in avian use of canopy gap, gap edge, and surrounding forest understory in a bottomland hardwood forest in the Upper Coastal Plain of South Carolina. We compared captures of foliage-gleaning birds among locations during four periods (spring migration, breeding, post-breeding, and fall migration). Foliage arthropod densities were greatest in the forest understory in all four seasons, but understory vegetation density was greatest in gaps. Foliage-gleaning bird abundance was positively associated with foliage-dwelling arthropods during the breeding (F  =  18.5, P < 0.001) and post-breeding periods (F  =  9.4, P  =  0.004), and negatively associated with foliage-dwelling arthropods during fall migration (F  =  5.4, P  =  0.03). Relationships between birds and arthropods were inconsistent, but the arthropod prey base seemed to be least important during migratory periods. Conversely, bird captures were positively correlated with understory vegetation density during all four periods (P < 0.001). Our study suggests high bird abundance associated with canopy gaps during the non-breeding period resulted less from high arthropod food resource availability than from complex understory and midstory vegetation structure.}, number={1}, journal={WILSON JOURNAL OF ORNITHOLOGY}, author={Moorman, Christopher E. and Bowen, Liessa T. and Kilgo, John C. and Hanula, James L. and Horn, Scott and Ulyshen, Michael D.}, year={2012}, month={Mar}, pages={31–39} } @article{golden_peterson_deperno_bardon_moorman_2012, title={Factors shaping private landowner engagement in wildlife management}, volume={37}, ISSN={1938-5463}, url={http://dx.doi.org/10.1002/wsb.235}, DOI={10.1002/wsb.235}, abstractNote={Abstract}, number={1}, journal={Wildlife Society Bulletin}, publisher={Wiley}, author={Golden, Katherine E. and Peterson, M. Nils and DePerno, Christopher S. and Bardon, Robert E. and Moorman, Christopher E.}, year={2012}, month={Dec}, pages={94–100} } @article{shake_moorman_riddle_burchell_2012, title={Influence of patch size and shape on occupancy by shrubland birds}, volume={14}, number={2}, journal={Condor}, author={Shake, C. S. and Moorman, C. E. and Riddle, J. D. and Burchell, M. R.}, year={2012}, pages={268–278} } @article{allen_moorman_peterson_hess_moore_2012, title={Overcoming socio-economic barriers to conservation subdivisions: A case-study of four successful communities}, volume={106}, ISSN={0169-2046}, url={http://dx.doi.org/10.1016/j.landurbplan.2012.03.012}, DOI={10.1016/j.landurbplan.2012.03.012}, abstractNote={Conservation subdivisions have emerged as an option to conserve open space, protect water quality and wildlife habitat, and maintain scenic views without compromising property rights. Despite economic and ecological advantages over traditional subdivisions, conservation subdivisions remain rare. We used a mixed-method study combining a survey of 246 people who attended conservation subdivision workshops with a qualitative case study of four communities that successfully developed conservation subdivisions to identify potential barriers to conservation subdivisions and strategies to overcome those barriers. A principal component analysis based on survey respondent rankings grouped barriers into: resistance to change among stakeholders, concerns about differences between traditional subdivisions and conservation subdivisions, concerns about consumer demand, and misperceptions about construction costs. Survey respondents indicated the top barrier to completion of conservation subdivisions was lack of incentives for developers. The case study communities overcame resistance from developers and landowners through educational efforts including informal meetings, charrettes, and workshops focusing on the environmental and economic benefits of conservation subdivisions. The communities had support from elected officials, and planning staff devoted necessary resources to rewrite ordinances, review sketch plans, and perform site visits. To overcome barriers to conservation subdivisions, communities could provide incentives including density bonuses and expedited approval processes. Encouraging participation in workshops and design charrettes for proposed developments may alleviate concerns of landowners who perceive a loss of property rights from new regulations and aid in the acceptance of conservation subdivisions.}, number={3}, journal={Landscape and Urban Planning}, publisher={Elsevier BV}, author={Allen, Stephen C. and Moorman, Christopher E. and Peterson, M. Nils and Hess, George R. and Moore, Susan E.}, year={2012}, month={Jun}, pages={244–252} } @article{raybuck_moorman_greenberg_deperno_gross_simon_warburton_2012, title={Short-term response of small mammals following oak regeneration silviculture treatments}, volume={274}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2012.02.012}, abstractNote={Upland, mixed-oak forests in the eastern United States have experienced widespread oak regeneration failure, largely due to cessation of anthropogenic disturbance. Silvicultural practices used to promote advance oak regeneration may affect ground-dwelling mammals. From May to August 2008 (pre-treatment), 2010 (first year post-treatment), and 2011 (second year post-treatment), we trapped small mammals to assess changes in species richness and abundance following experimental tests of three silvicultural treatments (prescribed burns, midstory herbicide applications, and shelterwood harvests) used to promote oak regeneration. We trapped small mammals in five replicates of each treatment and controls using Sherman live traps (2008 and 2010) and drift fences (2008, 2010, and 2011). From pre- to post-treatment, we evaluated the change in estimated peromyscid abundance and relative abundance of masked shrews (Sorex cinereus), smoky shrews (Sorex fumeus), and northern short-tailed shrews (Blarina brevicauda). Additionally, we evaluated the change in species richness across treatments for both sampling techniques. For all measures analyzed (i.e., species richness, peromyscid abundance, and relative abundance of shrews), the change from pre- to post-treatment did not differ among treatments. However, more masked shrews, smoky shrews, and northern short-tailed shrews were captured in 2011 (i.e., second year post-treatment) than in 2010 (i.e., first year post-treatment). Our research indicates that, in the short-term, small mammals (e.g., mice and shrews) can tolerate a wide range of forest disturbance following oak regeneration treatments. However, delayed treatment effects (e.g., additional post-herbicide midstory dieback) or additive changes following future treatments (e.g., prescribed burns following shelterwood harvests or multiple prescribed burns) may compound effects on small mammal populations, and should be assessed with long-term research (>2 years post-treatment).}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Raybuck, Amy L. and Moorman, Christopher E. and Greenberg, Cathryn H. and DePerno, Christopher S. and Gross, Kevin and Simon, Dean M. and Warburton, Gordon S.}, year={2012}, month={Jun}, pages={10–16} } @article{moorman_plush_orr_reberg-horton_gardner_2012, title={Small mammal use of field borders planted as beneficial insect habitat}, volume={37}, ISSN={1938-5463}, url={http://dx.doi.org/10.1002/wsb.226}, DOI={10.1002/wsb.226}, abstractNote={Abstract}, number={1}, journal={Wildlife Society Bulletin}, publisher={Wiley}, author={Moorman, Christopher E. and Plush, Charles J. and Orr, David B. and Reberg-Horton, Chris and Gardner, Beth}, year={2012}, month={Nov}, pages={209–215} } @article{shake_moorman_burchell_2011, title={Cropland Edge, Forest Succession, and Landscape Affect Shrubland Bird Nest Predation}, volume={75}, ISSN={["0022-541X"]}, DOI={10.1002/jwmg.101}, abstractNote={Abstract}, number={4}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Shake, Corey S. and Moorman, Christopher E. and Burchell, Michael R., II}, year={2011}, month={May}, pages={825–835} } @article{dellinger_mcvey_cobb_moorman_2011, title={Diameter thresholds for distinguishing between red wolf and other canid scat}, volume={35}, ISSN={1938-5463}, url={http://dx.doi.org/10.1002/wsb.60}, DOI={10.1002/wsb.60}, abstractNote={Abstract}, number={4}, journal={Wildlife Society Bulletin}, publisher={Wiley}, author={Dellinger, Justin A. and Mcvey, Justin M. and Cobb, David T. and Moorman, Christopher E.}, year={2011}, month={Sep}, pages={416–420} } @article{blackman_deperno_heiniger_krachey_moorman_peterson_2011, title={Effects of crop field characteristics on nocturnal winter use by American woodcock}, volume={76}, ISSN={0022-541X}, url={http://dx.doi.org/10.1002/JWMG.254}, DOI={10.1002/jwmg.254}, abstractNote={Abstract}, number={3}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Blackman, Emily B. and Deperno, Christopher S. and Heiniger, Ron W. and Krachey, Matthew J. and Moorman, Christopher E. and Peterson, M. Nils}, year={2011}, month={Nov}, pages={528–533} } @inproceedings{golden_deperno_moorman_peterson_bardon_2011, title={Predicting North Carolina landowner participation and interest in wildlife related fee access}, booktitle={Proceedings of the Annual Conference of the Southeastern Association of Fish and Wildlife Agencies}, author={Golden, K. E. and DePerno, C. S. and Moorman, C. E. and Peterson, N. and Bardon, R. E.}, year={2011}, pages={21–26} } @article{ayers_moorman_deperno_yelverton_wang_2010, title={Effects of Mowing on Anthraquinone for Deterrence of Canada Geese}, volume={74}, ISSN={["0022-541X"]}, DOI={10.2193/2009-323}, abstractNote={ABSTRACT Anthraquinone (AQ)‐based repellents have been shown to reduce Canada goose (Branta canadensis) use of turfgrass; however, impacts of frequent mowing on efficacy of AQ have not been studied. Our objective was to determine efficacy and longevity of a rain‐fast AQ‐based avian repellent, FlightControl® PLUS (FCP), as a deterrent of free‐ranging resident Canada geese under 2 mowing frequencies. We conducted the study at 8 sites in the Triangle region (Raleigh, Durham, and Chapel Hill) of North Carolina, USA. We arranged our experiment in a randomized complete block design, with each of 8 sites containing 4 0.1‐ha treatment combinations: 1) treated with FCP and mowed every 4 days (T4), 2) treated with FCP and mowed every 8 days (T8), 3) untreated and mowed every 4 days, and 4) untreated and mowed every 8 days. We conducted 4 37‐day field sessions (Jun‐Jul 2007, Sep‐Oct 2007, Jun‐Jul 2008, and Sep‐Oct 2008), representing the summer molting phase and the full‐plumage phase. Resident goose use (measured by daily no. of droppings) was 41–70% lower on treated plots than on untreated plots, but use was similar between T4 and T8. Average FCP coverage on grass blades decreased in coverage from approximately 95% to 10% over the 30‐day posttreatment phase. Results indicate that resident Canada goose use of FCP‐treated turfgrass areas was lower than untreated areas even when chemical coverage on grass was 10%. Further, mowing frequency did not have a clear impact on the efficacy of FCP as a Canada goose repellent.}, number={8}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Ayers, Christopher R. and Moorman, Christopher E. and Deperno, Christopher S. and Yelverton, Fred H. and Wang, Huixia J.}, year={2010}, month={Nov}, pages={1863–1868} } @article{kleist_moorman_deperno_bardon_2010, title={Opportunities for planned county-based wildlife programming}, volume={48}, number={2}, journal={Journal of Extension}, author={Kleist, A. M. and Moorman, C. E. and DePerno, C. S. and Bardon, R. E.}, year={2010} } @article{savage_moorman_gerwin_sorenson_2010, title={PREY SELECTION BY SWAINSON'S WARBLERS ON THE BREEDING GROUNDS}, volume={112}, ISSN={["1938-5129"]}, DOI={10.1525/cond.2010.090055}, abstractNote={Abstract. Swainson's Warbler (Limnothlypis swainsonii) breeds in bottomland hardwood forests across the southeastern United States, where it is believed to be one of the rarest breeding songbirds. Although information on its nest-site habitat is considerable, little is known about its foraging habitat except that the species is insectivorous, with a large bill used to flip fallen leaves on the forest floor. We captured Swainson's Warblers and flushed their crops to determine their diet and sampled leaf-litter arthropods and vegetation at each location of capture. We compared the proportion of arthropod orders in the crop samples to the proportion of arthropods collected in the leaf litter to determine the warbler's prey in proportion to its availability. Although Acari (mites and ticks) and Chilopoda (centipedes) were the most abundant arthropods in the leaf-litter samples (51% and 18%, respectively), these orders rarely occurred in the warblers' crops. Conversely, Araneae (spiders) and Coleoptera (beetles) were uncommon in leaf-litter samples (2% and 5%, respectively) but were the most abundant arthropod orders in the warblers' crops. Binary logistic regression with presence or absence of Araneae as the response variable and habitat measures as the predictor variables revealed that the probability of spiders occurring in the leaf litter increased as leaf-litter depth increased. To promote foraging habitat for Swainson's Warbler, deep leaf litter should be maintained by maintaining patches of closed-canopy forests and restoring natural regimes of flooding.}, number={3}, journal={CONDOR}, author={Savage, Amelia L. and Moorman, Christopher E. and Gerwin, John A. and Sorenson, Clyde}, year={2010}, month={Aug}, pages={605–614} } @article{matthews_moorman_greenberg_waldrop_2010, title={Response of reptiles and amphibians to repeated fuel reduction treatments}, volume={74}, DOI={10.1111/j.1937-2817.2010.tb01251.x}, abstractNote={Abstract:Recent use of prescribed fire and fire surrogates to reduce fuel hazards has spurred interest in their effects on wildlife. Studies of fire in the southern Appalachian Mountains (USA) have documented few effects on reptiles and amphibians. However, these studies were conducted after only one fire and for only a short time (1–3 yr) after the fire. From mid‐May to mid‐August 2006 and 2007, we used drift fences with pitfall and funnel traps to capture reptiles and amphibians in a control and 3 replicated fuel‐reduction treatments: 1) twice‐burned (2003 and 2006), 2) mechanical understory cut (2002), and 3) mechanical understory cut (2002) followed by 2 burns (2003 and 2006). We captured fewer salamanders in mechanical + twice‐burned treatment areas than in twice‐burned and control treatment areas, but we captured more lizards in mechanical + twice‐burned treatment areas than in other treatment areas. Higher lizard captures in mechanical + twice‐burned treatment areas likely was related to increased ground temperatures and greater thermoregulatory opportunities. Higher and more variable ground temperatures and faster drying of remaining litter and duff may have led to fewer salamander captures in mechanical + twice‐burned treatment areas. Our longer term results, after 2 prescribed burns, differ from shorter term results. After one prescribed burn at the same site, eastern fence lizard (Sceloporus undulatus) captures were greater in mechanical + burn treatment areas but salamander captures did not differ among treatment areas. Our results indicate that multiple (≥2) fuel‐reduction treatments that decrease canopy cover may benefit lizards but negatively affect salamanders.}, number={6}, journal={Journal of Wildlife Management}, author={Matthews, C. E. and Moorman, C. E. and Greenberg, C. H. and Waldrop, T. A.}, year={2010}, pages={1301–1310} } @article{riddle_stanislav_pollock_moorman_perkins_2010, title={Separating Components of the Detection Process With Combined Methods: An Example With Northern Bobwhite}, volume={74}, ISSN={["1937-2817"]}, DOI={10.2193/2009-220}, abstractNote={Abstract There are various methods of estimating detection probabilities for avian point counts. Distance and multiple-observer methods require the sometimes unlikely assumption that all birds in the population are available (i.e., sing or are visible) during a count, but the time-of-detection method allows for the possibility that some birds are unavailable during the count. We combined the dependent double-observer method with the time-of-detection method and obtained field-based estimates of the components of detection probability for northern bobwhite (Colinus virginianus). Our approach was a special case of Pollock's robust capture–recapture design where the probability that a bird does not sing is analogous to the probability that an animal is a temporary emigrant. Top models indicated that observers' detection probabilities were similar (0.78–0.84) if bobwhite were available, but bobwhite only had an approximately 0.61 probability of being available during a 2.5-minute sampling interval. Additionally, observers' detection probabilities increased substantially after the initial encounter with an individual bobwhite (analogous to a trap-happy response on the part of the observer). A simulated data set revealed that the combined method was precise when availability and detection given availability were substantially lower. Combined methods approaches can provide critical information for researchers and land managers to make decisions regarding survey length and personnel requirements for point-count–based surveys.}, number={6}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Riddle, Jason D. and Stanislav, Stephen J. and Pollock, Kenneth H. and Moorman, Christopher E. and Perkins, Fern S.}, year={2010}, month={Aug}, pages={1319–1325} } @article{riddle_moorman_2010, title={The importance of agriculture-dominated landscapes and lack of field border effect for early-succession songbird nest success}, volume={5}, DOI={10.5751/ace-00424-050209}, abstractNote={In recent decades, many early-succession songbird species have experienced severe and widespread declines, which often are related to habitat destruction. Field borders create additional or enhance existing early-succession habitat on farmland. However, field border shape and the landscape context surrounding farms may influence the effectiveness of field borders in contributing to the stabilization or increase of early-succession bird populations. We examined the influence of linear and nonlinear field borders on farms in landscapes dominated by either agriculture or forests on nest success and Brown-headed Cowbird (Molothrus ater) brood parasitism of Indigo Bunting (Passerina cyanea) and Blue Grosbeak (Passerina caerulea) nests combined. Field border establishment did not affect nest survival probability and brood parasitism frequency of Indigo Bunting and Blue Grosbeak nests. Indigo Bunting/Blue Grosbeak nest success probability was more than twice as high in agriculture-dominated landscapes (39%) than in forested landscapes (17%), and brood parasitism frequency was high (33%) but did not differ between landscapes. Edges in agriculture-dominated landscapes can be higher-quality habitats for early-succession birds than edges in forest-dominated landscapes, but our field border treatments did not enhance nest success for these birds on farms in either landscape. RESUME. Au cours des dernieres decennies, de nombreux passereaux de debut de succession ont subi un declin marque et generalise, souvent lie a la destruction d’habitat. Les lisieres de champs representent de nouveaux milieux de debut de succession ou s’ajoutent aux milieux de debut de succession deja existants dans les paysages agricoles. Toutefois, la configuration des lisieres et le contexte paysager aux environs des fermes peuvent avoir une influence sur l’efficacite des lisieres, en contribuant a la stabilisation ou a l’augmentation des populations d’oiseaux de debut de succession. Nous avons examine l’effet de lisieres lineaires et non lineaires sur des fermes situees dans des paysages a dominance agricole ou forestiere, sur deux parametres du Passerin indigo (Passerina cyanea) et du Guiraca bleu (Passerina caerulea) : leur succes de nidification et leur taux de parasitisme par le Vacher a tete brune (Molothrus ater). La creation de lisieres n’a pas eu d’effets sur le taux de survie des nids, ni sur le taux de parasitisme. Le taux de survie des nids de Passerin indigo et de Guiraca bleu dans les paysages agricoles (39 %) etait plus du double de celui observe dans les paysages forestiers (17 %). Le taux de parasitisme etait eleve (33 %), mais ne differait pas entre les deux types de paysages. Les lisieres presentes dans les paysages domines par l’agriculture s’averent etre des milieux de meilleure qualite pour les oiseaux de debut de succession que ne le sont les University of Wisconsin Stevens Point, North Carolina State University Avian Conservation and Ecology 5(2): 9 http://www.ace-eco.org/vol5/iss2/art9/ lisieres presentes dans les paysages domines par les forets. Les traitements que nous avons effectues sur les lisieres n’ont toutefois pas permis d’augmenter le succes de nidification de ces oiseaux dans l’un ou l’autre des deux types de paysages.}, number={2}, journal={Avian Conservation and Ecology}, author={Riddle, J. D. and Moorman, C. E.}, year={2010} } @article{champlin_kilgo_gumpertz_moorman_2009, title={Avian Response to Microclimate in Canopy Gaps in a Bottomland Hardwood Forest}, volume={8}, ISSN={["1938-5412"]}, DOI={10.1656/058.008.0110}, abstractNote={Abstract Microclimate may influence use of early successional habitat by birds. We assessed the relationships between avian habitat use and microclimate (temperature, light intensity, and relative humidity) in experimentally created canopy gaps in a bottomland hardwood forest on the Savannah River Site, SC. Gaps were 2- to 3-year-old group-selection timber harvest openings of three sizes (0.13, 0.26, 0.50 ha). Our study was conducted from spring through fall, encompassing four bird-use periods (spring migration, breeding, post-breeding, and fall migration), in 2002 and 2003. We used mist netting and simultaneously recorded microclimate variables to determine the influence of microclimate on bird habitat use. Microclimate was strongly affected by net location within canopy gaps in both years. Temperature generally was higher on the west side of gaps, light intensity was greater in gap centers, and relative humidity was higher on the east side of gaps. However, we found few relationships between bird captures and the microclimate variables. Bird captures were inversely correlated with temperature during the breeding and post-breeding periods in 2002 and positively correlated with temperature during spring 2003. Captures were high where humidity was high during post-breeding 2002, and captures were low where humidity was high during spring 2003. We conclude that variations in the local microclimate had minor influence on avian habitat use within gaps. Instead, habitat selection in relatively mild regions like the southeastern US is based primarily on vegetation structure, while other factors, including microclimate, are less important.}, number={1}, journal={SOUTHEASTERN NATURALIST}, author={Champlin, Tracey B. and Kilgo, John C. and Gumpertz, Marcia L. and Moorman, Christopher E.}, year={2009}, pages={107–120} } @article{kohut_hess_moorman_2009, title={Avian use of suburban greenways as stopover habitat}, volume={12}, DOI={10.1007/s11252-009-0099-6}, abstractNote={Greenways may provide stopover habitat for migrating birds in otherwise inhospitable suburban landscapes. We examined the effect of greenway forested corridor width, vegetation composition and structure, and adjacent land cover on the species richness and abundance of migrating songbirds during spring and fall migration in Raleigh and Cary, North Carolina, USA. Generally, migrating birds were more abundant in wider forest corridors during spring and fall migration. During the spring, migrants were detected more commonly in greenways with taller trees and a higher percentage of hardwood trees. In the fall, migrant richness and abundance was highest in greenways with lower canopy cover, possibly because of the increased vertical complexity of the vegetation at these sites. Forest-interior migrant richness was not correlated with corridor width in either season, but these species were more abundant in greenways bordered by less bare earth and pavement cover in the spring. No other bird groupings were correlated with adjacent land cover measures. Although migrants used greenways of all widths, forested corridors wider than 150 m should be conserved whenever possible to provide stopover habitat for forest-interior migrants. Shrub cover should be retained to maintain vegetative complexity. Habitat for the greatest diversity of migrants can be provided by constructing greenways in areas of lower development intensity and encouraging residents to retain shrubs and trees on properties bordering greenways.}, number={4}, journal={Urban Ecosystems}, author={Kohut, S. and Hess, G. and Moorman, C.}, year={2009}, pages={487–502} } @article{champlin_kilgo_moorman_2009, title={Food abundance does not determine bird use of early-successional habitat}, volume={90}, ISSN={["1939-9170"]}, DOI={10.1890/08-1190.1}, abstractNote={Few attempts have been made to experimentally address the extent to which temporal or spatial variation in food availability influences avian habitat use. We used an experimental approach to investigate whether bird use differed between treated (arthropods reduced through insecticide application) and control (untreated) forest canopy gaps within a bottomland hardwood forest in the Upper Coastal Plain of South Carolina, USA. Gaps were two‐ to three‐year‐old group selection timber harvest openings of three sizes (0.13, 0.26, and 0.50 ha). Our study was conducted during four bird use periods (spring migration, breeding, post‐breeding, and fall migration) in 2002 and 2003. Arthropods were reduced in treated gaps by 68% in 2002 and 73% in 2003. We used mist‐netting captures and foraging attack rates to assess the influence of arthropod abundance on avian habitat use. Evidence that birds responded to arthropod abundance was limited and inconsistent. In 2002, we generally captured more birds in treated gaps of the smallest size (0.13 ha) and fewer birds in treated gaps of the larger sizes. In 2003, we recorded few differences in the number of captures in treated and control gaps. Foraging attack rates generally were lower in treated than in control gaps, indicating that birds were able to adapt to the reduced food availability and remain in treated gaps. We conclude that arthropod abundance was not a proximate factor controlling whether forest birds used our gaps. The abundance of food resources may not be as important in determining avian habitat selection as previous research has indicated, at least for passerines in temperate subtropical regions.}, number={6}, journal={ECOLOGY}, author={Champlin, Tracey B. and Kilgo, John C. and Moorman, Christopher E.}, year={2009}, month={Jun}, pages={1586–1594} } @article{matthews_moorman_greenberg_waldrop_2009, title={Response of soricid populations to repeated fire and fuel reduction treatments in the southern Appalachian Mountains}, volume={257}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2009.02.006}, abstractNote={Fuel hazards have increased in forests across the United States because of fire exclusion during the 20th century. Treatments used to reduce fuel buildup may affect wildlife, such as shrews, living on the forest floor, especially when treatments are applied repeatedly. From mid-May to mid-August 2006 and 2007, we used drift fences with pitfall traps to capture shrews in western North Carolina in 3 fuel reduction treatment areas [(1) twice-burned (2003 and 2006), (2) mechanical understory cut (2002), and (3) mechanical understory cut (2002) followed by 2 burns (2003 and 2006)] and a control. We captured 77% fewer southeastern shrews (Sorex longirostris) in mechanical + twice-burned treatment areas than in mechanical treatment areas in 2006, but southeastern shrew captures did not differ among treatment areas in 2007. Total shrew captures did not differ among treatment areas in either year. Decreases in leaf litter, duff depth, and canopy cover in mechanical + twice-burned treatment areas may have decreased ground-level moisture, thereby causing short-term declines in southeastern shrew captures. Prescribed fire or mechanical fuel reduction treatments in the southern Appalachian Mountains did not greatly affect shrew populations, though the combination of both treatments may negatively affect some shrew species, at least temporarily.}, number={9}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Matthews, Charlotte E. and Moorman, Christopher E. and Greenberg, Cathryn H. and Waldrop, Thomas A.}, year={2009}, month={Apr}, pages={1939–1944} } @article{riddle_moorman_pollock_2008, title={A comparison of methods for estimating northern bobwhite covey detection probabilities}, volume={72}, DOI={10.2193/2007-435}, abstractNote={Abstract: We compared the time‐of‐detection and logistic regression methods of estimating probability of detection for northern bobwhite (Colinus virginianus) coveys. Both methods are unusual in that they allow estimation of the total probability of detection (i.e., the product of the probability that a covey is available for detection [i.e., that a covey vocalizes] and detection given availability). The logistic regression method produced an average detection probability of 0.596 (SE = 0.020) and the time‐of‐detection method produced a detection probability estimate of 0.540 (SE = 0.086), and the 2 estimates were not significantly different. This is the first evaluation of the time‐of‐detection method with empirical field data. Although the time‐of‐detection and logistic regression method each have advantages, both can be used under appropriate conditions to improve estimates of bobwhite abundance by allowing for the estimation of detection probabilities. Improved estimates of bobwhite abundance will allow land managers to make more informed management decisions.}, number={6}, journal={Journal of Wildlife Management}, author={Riddle, J. D. and Moorman, C. E. and Pollock, K. H.}, year={2008}, pages={1437–1442} } @article{smith_osmond_moorman_stucky_gilliam_2008, title={Effect of vegetation management on bird habitat in Riparian buffer zones}, volume={7}, ISSN={["1938-5412"]}, DOI={10.1656/1528-7092(2008)7[277:EOVMOB]2.0.CO;2}, abstractNote={Abstract Riparian buffers can be valuable refuge areas for wildlife in otherwise homogeneous agricultural landscapes. Government sponsored programs like the Cropland Reserve Program generally require the planting of specific vegetative species during buffer restoration, although the effectiveness of such an approach when compared to restoration by volunteer species is unknown. We studied the effect of differences in vegetation structure on avian habitat in riparian buffer zones. A 25 m (82 ft) wide planted woodland buffer, 30 m (98 ft) wide grass, shrub, and woodland three-zone buffer, and a 9 m (30 ft) wide shrub buffer were evaluated for habitat potential using breeding-bird counts and vegetation surveys. Bird density and species richness varied with the structure of the vegetative communities present at the three sites. Avian species richness and total detections were higher in the three-zone buffer than in both the shrub and planted buffer, likely a result of the diversity of vegetation at the site. These data suggest that restoration of riparian areas by allowing fallow vegetation to recolonize is at the very least equally beneficial to avian wildlife as is restoration by planting specific grass, shrub, and tree species. Buffer restoration by natural revegetation using this method could be recommended as an alternative to implementation by planting riparian species due to its simplicity and cost effectiveness.}, number={2}, journal={SOUTHEASTERN NATURALIST}, author={Smith, Timothy A. and Osmond, Deanna L. and Moorman, Christopher E. and Stucky, Jon M. and Gilliam, J. Wendell}, year={2008}, pages={277–288} } @article{riddle_moorman_pollock_2008, title={The importance of habitat shape and landscape context to northern bobwhite populations}, volume={72}, ISSN={["1937-2817"]}, DOI={10.2193/2007-469}, abstractNote={Abstract: Northern bobwhite (Colinus virginianus) populations have declined nationally for at least the past 4 decades. Field borders have been promoted as an important component of conservation plans to reverse this decline. Field border characteristics, such as shape and the landscapes in which the borders are established, have the potential to influence their effectiveness for recovering northern bobwhite populations. We established narrow linear (approx. 3‐m‐wide) and nonlinear field borders on farms in agriculture‐dominated and forest‐dominated landscapes in the Coastal Plain of North Carolina, USA, after collecting pretreatment data on summer bobwhite abundance. After establishment of field borders, summer bobwhite abundance nearly doubled on farms in agriculture‐dominated landscapes and increased approximately 57% on farms with nonlinear field borders. Summer bobwhite abundance did not increase on farms with linear field borders in forest‐dominated landscapes. Nonlinear and narrow linear field borders can be used to increase bobwhite numbers on farms in landscapes dominated by agriculture. Less flexibility exists in forest‐dominated landscapes, where we found only nonlinear field borders resulted in an increase.}, number={6}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Riddle, Jason D. and Moorman, Christopher E. and Pollock, Kenneth H.}, year={2008}, month={Aug}, pages={1376–1382} } @article{mason_moorman_hess_sinclair_2007, title={Designing suburban greenways to provide habitat for forest-breeding birds}, volume={80}, ISSN={["1872-6062"]}, DOI={10.1016/j.landurbplan.2006.07.002}, abstractNote={Appropriately designed, greenways may provide habitat for neotropical migrants, insectivores, and forest-interior specialist birds that decrease in diversity and abundance as a result of suburban development. We investigated the effects of width of the forested corridor containing a greenway, adjacent land use and cover, and the composition and vegetation structure within the greenway on breeding bird abundance and community composition in suburban greenways in Raleigh and Cary, North Carolina, USA. Using 50 m fixed-radius point counts, we surveyed breeding bird communities for 2 years at 34 study sites, located at the center of 300-m-long greenway segments. Percent coverage of managed area within the greenway, such as trail and other mowed or maintained surfaces, was a predictor for all development-sensitive bird groupings. Abundance and richness of development-sensitive species were lowest in greenway segments containing more managed area. Richness and abundance of development-sensitive species also decreased as percent cover of pavement and bare earth adjacent to greenways increased. Urban adaptors and edge-dwelling birds, such as Mourning Dove, House Wren, House Finch, and European Starling, were most common in greenways less than 100 m wide. Conversely, forest-interior species were not recorded in greenways narrower than 50 m. Some forest-interior species, such as Acadian Flycatcher, Hairy Woodpecker, and Wood Thrush, were recorded primarily in greenways wider than 100 m. Others, including ground nesters such as Black-and-white Warbler, Louisiana Waterthrush, and Ovenbird, were recorded only in greenways wider than 300 m. Landscape and urban planners can facilitate conservation of development-sensitive birds in greenways by minimizing the width of the trail and associated mowed and landscaped surfaces adjacent to the trail, locating trails near the edge of greenway forest corridors, and giving priority to the protection of greenway corridors at least 100 m wide with low levels of impervious surface (pavement, buildings) and bare earth in the adjacent landscape.}, number={1-2}, journal={LANDSCAPE AND URBAN PLANNING}, author={Mason, Jamie and Moorman, Christopher and Hess, George and Sinclair, Kristen}, year={2007}, month={Mar}, pages={153–164} } @article{bowen_moorman_kilgo_2007, title={Seasonal bird use of canopy gaps in a bottomland forest}, volume={119}, ISSN={["1938-5447"]}, DOI={10.1676/05-091.1}, abstractNote={Abstract Bird use of small canopy gaps within mature forests has not been well studied, particularly across multiple seasons. We investigated seasonal differences in bird use of gap and forest habitat within a bottomland hardwood forest in the Upper Coastal Plain of South Carolina. Gaps were 0.13- to 0.5-ha, 7- to 8-year-old group-selection timber harvest openings. Our study occurred during four bird-use periods (spring migration, breeding, postbreeding, and fall migration) in 2001 and 2002. We used plot counts and mist netting to estimate bird abundance in canopy gaps and surrounding mature forest habitats. Using both survey methods, we observed more birds, including forest-interior species, forest-edge species, field-edge species, and several individual species in canopy gap and gap-edge habitats than in surrounding mature forest during all periods. Interactions between period and habitat type often were significant in models, suggesting a seasonal shift in habitat use. Bird activity generally shifted between the interior of canopy gaps and the immediate gap edge, but many species increased their use of forested habitat during the breeding period. This suggests that many species of birds selectively choose gap and gap-edge habitat over surrounding mature forest during the non-breeding period. Creation of small canopy gaps within a mature forest may increase local bird species richness. The reasons for increased bird activity in gaps remain unclear.}, number={1}, journal={WILSON JOURNAL OF ORNITHOLOGY}, author={Bowen, Liessa T. and Moorman, Christopher E. and Kilgo, John C.}, year={2007}, month={Mar}, pages={77–88} } @article{moorman_bowen_kilgo_sorenson_hanula_horn_ulyshen_2007, title={Seasonal diets of insectivorous birds using canopy gaps in a bottomland forest}, volume={78}, ISSN={["1557-9263"]}, DOI={10.1111/j.1557-9263.2006.00081.x}, abstractNote={Little is known about how insectivorous bird diets are influenced by arthropod availability and about how these relationships vary seasonally. We captured birds in forest-canopy gaps and adjacent mature forest during 2001 and 2002 at the Savannah River Site in Barnwell County, South Carolina, and flushed their crops to gather information about arthropods eaten during four periods: spring migration, breeding, postbreeding, and fall migration. Arthropod availability for foliage- and ground-gleaning birds was examined by leaf clipping and pitfall trapping. Coleopterans and Hemipterans were used by foliage- and ground-gleaners more than expected during all periods, whereas arthropods in the orders Araneae and Hymenoptera were used as, or less than, expected based on availability during all periods. Ground-gleaning birds used Homopterans and Lepidopterans in proportions higher than availability during all periods. Arthropod use by birds was consistent from spring through fall migration, with no apparent seasonal shift in diet. Based on concurrent studies, heavily used orders of arthropods were equally abundant or slightly less abundant in canopy gaps than in the surrounding mature forest, but bird species were most frequently detected in gaps. Such results suggest that preferential feeding on arthropods by foliage-gleaning birds in gap habitats reduced arthropod densities or, alternatively, that bird use of gap and forest habitat was not determined by food resources. The abundance of arthropods across the stand may have allowed birds to remain in the densely vegetated gaps where thick cover provides protection from predators.}, number={1}, journal={JOURNAL OF FIELD ORNITHOLOGY}, author={Moorman, Christopher E. and Bowen, Liessa T. and Kilgo, John C. and Sorenson, Clyde E. and Hanula, James L. and Horn, Scott and Ulyshen, Mike D.}, year={2007}, pages={11–20} } @article{miller_hess_moorman_2006, title={Southern two-lined salamanders in urbanizing watersheds}, volume={10}, ISSN={1083-8155 1573-1642}, url={http://dx.doi.org/10.1007/s11252-006-0012-5}, DOI={10.1007/s11252-006-0012-5}, number={1}, journal={Urban Ecosystems}, publisher={Springer Science and Business Media LLC}, author={Miller, Jennifer E. and Hess, George R. and Moorman, Christopher E.}, year={2006}, month={Dec}, pages={73–85} } @article{ulyshen_hanula_horn_kilgo_moorman_2006, title={The response of ground beetles (Coleoptera : Carabidae) to selection cutting in a South Carolina bottomland hardwood forest}, volume={15}, ISSN={["1572-9710"]}, DOI={10.1007/s10531-004-6899-3}, abstractNote={We compared the response of ground beetles (Coleoptera: Carabidae) to the creation of canopy gaps of different size (0.13, 0.26, and 0.50 ha) and age (1 and 7 years) in a bottomland hardwood forest (South Carolina, USA). Samples were collected four times in 2001 by malaise and pitfall traps placed at the center and edge of each gap, and 50 m into the surrounding forest. Species richness was higher at the center of young gaps than in old gaps or in the forest, but there was no statistical difference in species richness between old gaps and the forests surrounding them. Carabid abundance followed the same trend, but only with the exclusion of Semiardistomis viridis (Say), a very abundant species that differed in its response to gap age compared to most other species. The carabid assemblage at the gap edge was very similar to that of the forest, and there appeared to be no distinct edge community. Species known to occur in open or disturbed habitats were more abundant at the center of young gaps than at any other location. Generalist species were relatively unaffected by the disturbance, but one species (Dicaelus dilatatus Say) was significantly less abundant at the centers of young gaps. Forest inhabiting species were less abundant at the centers of old gaps than in the forest, but not in the centers of young gaps. Comparison of community similarity at various trapping locations showed that communities at the centers of old and young gaps had the lowest similarity (46.5%). The community similarity between young gap centers and nearby forest (49.1%) and old gap centers and nearby forest (50.0%) was similarly low. These results show that while the abundance and richness of carabids in old gaps was similar to that of the surrounding forest, the species composition between the two sites differed greatly.}, number={1}, journal={BIODIVERSITY AND CONSERVATION}, author={Ulyshen, MD and Hanula, JL and Horn, S and Kilgo, JC and Moorman, CE}, year={2006}, month={Jan}, pages={261–274} } @article{ulyshen_hanula_horn_kilgo_moorman_2005, title={Herbivorous insect response to group selection cutting in a southeastern bottomland hardwood forest}, volume={34}, ISSN={["1938-2936"]}, DOI={10.1603/0046-225X-34.2.395}, abstractNote={Abstract Malaise and pitfall traps were used to sample herbivorous insects in canopy gaps created by group-selection cutting in a bottomland hardwood forest in South Carolina. The traps were placed at the centers, edges, and in the forest adjacent to gaps of different sizes (0.13, 0.26, and 0.50 ha) and ages (1 and 7 yr old) during four sampling periods in 2001. Overall, the abundance and species richness of insect herbivores were greater at the centers of young gaps than at the edge of young gaps or in the forest surrounding young gaps. There were no differences in abundance or species richness among old gap locations (i.e., centers, edges, and forest), and we collected significantly more insects in young gaps than old gaps. The insect communities in old gaps were more similar to the forests surrounding them than young gap communities were to their respective forest locations, but the insect communities in the two forests locations (surrounding young and old gaps) had the highest percent similarity of all. Although both abundance and richness increased in the centers of young gaps with increasing gap size, these differences were not significant. We attribute the increased numbers of herbivorous insects to the greater abundance of herbaceous plants available in young gaps.}, number={2}, journal={ENVIRONMENTAL ENTOMOLOGY}, author={Ulyshen, MD and Hanula, JL and Horn, S and Kilgo, JC and Moorman, CE}, year={2005}, month={Apr}, pages={395–402} } @article{sinclair_hess_moorman_mason_2005, title={Mammalian nest predators respond to greenway width, landscape context and habitat structure}, volume={71}, ISSN={["0169-2046"]}, DOI={10.1016/j.landurbplan.2004.04.001}, abstractNote={Birds of conservation concern breed in suburban greenways, yet abundant populations of mammals that depredate bird nests might reduce nest success. We evaluated how three factors influenced the abundance of mammalian nest predators in thirty-four 300-m long forested greenway segments in Raleigh and Cary, North Carolina, USA: (1) the width of the forested corridor containing the greenway, (2) the land-use adjacent to the forested corridor, and (3) the habitat structure within the greenway. Forest corridor width and adjacent land-use were measured using aerial photographs. Attributes of adjacent land use included categorical measures of development intensity (low-density residential, high-density residential, office/institutional), and the proportions of forest canopy, grass, buildings, and pavement. Several measures of habitat structure within the greenway were collected in the field, including trail width and surface type, and percentage of mature forest. We measured the relative abundance of mammalian nest predators with scent-station transects, operated for five nights during the 2002 breeding bird season. Total abundance of mammalian nest predators increased significantly as forest corridor width decreased. We found no relationship between categorical measures of land-use and total abundance of mammalian nest predators. Specific attributes of the landscape adjacent to the greenway, however, did have an effect. Greenways adjacent to landscapes with fewer buildings had a higher abundance of mammalian nest predators. The abundance of individual species varied with the amount of canopy, lawn, and pavement in the adjacent landscape. Some measures of habitat structure of greenways also were correlated with the abundance of mammalian nest predators. Greenway segments with wider trails had a higher abundance of mammalian nest predators, as did segments with a higher percentage of mature forest. No habitat structure variables were significant for all species. To reduce the overall risk of avian nest predation by mammals, forested greenways should be designed with wider forest corridors and narrower, unpaved trails. Some greenway characteristics that favor high-nest predator populations also favor birds of conservation concern. Similarly, some characteristics correlated with lower predator occurrence are also correlated with lower abundance of birds of conservation concern. Thus, management of greenways and the surrounding landscape must balance reduction of predator communities with the promotion of desired bird communities and other conservation goals.}, number={2-4}, journal={LANDSCAPE AND URBAN PLANNING}, author={Sinclair, KE and Hess, GR and Moorman, CE and Mason, JH}, year={2005}, month={Mar}, pages={277–293} } @article{sinclair_hess_moorman_mason_2005, title={Mammalian nest predators respond to greenway width, landscape context and habitat structure}, volume={71}, ISSN={0169-2046}, url={http://dx.doi.org/10.1016/S0169-2046(04)00082-9}, DOI={10.1016/S0169-2046(04)00082-9}, abstractNote={Birds of conservation concern breed in suburban greenways, yet abundant populations of mammals that depredate bird nests might reduce nest success. We evaluated how three factors influenced the abundance of mammalian nest predators in thirty-four 300-m long forested greenway segments in Raleigh and Cary, North Carolina, USA: (1) the width of the forested corridor containing the greenway, (2) the land-use adjacent to the forested corridor, and (3) the habitat structure within the greenway. Forest corridor width and adjacent land-use were measured using aerial photographs. Attributes of adjacent land use included categorical measures of development intensity (low-density residential, high-density residential, office/institutional), and the proportions of forest canopy, grass, buildings, and pavement. Several measures of habitat structure within the greenway were collected in the field, including trail width and surface type, and percentage of mature forest. We measured the relative abundance of mammalian nest predators with scent-station transects, operated for five nights during the 2002 breeding bird season. Total abundance of mammalian nest predators increased significantly as forest corridor width decreased. We found no relationship between categorical measures of land-use and total abundance of mammalian nest predators. Specific attributes of the landscape adjacent to the greenway, however, did have an effect. Greenways adjacent to landscapes with fewer buildings had a higher abundance of mammalian nest predators. The abundance of individual species varied with the amount of canopy, lawn, and pavement in the adjacent landscape. Some measures of habitat structure of greenways also were correlated with the abundance of mammalian nest predators. Greenway segments with wider trails had a higher abundance of mammalian nest predators, as did segments with a higher percentage of mature forest. No habitat structure variables were significant for all species. To reduce the overall risk of avian nest predation by mammals, forested greenways should be designed with wider forest corridors and narrower, unpaved trails. Some greenway characteristics that favor high-nest predator populations also favor birds of conservation concern. Similarly, some characteristics correlated with lower predator occurrence are also correlated with lower abundance of birds of conservation concern. Thus, management of greenways and the surrounding landscape must balance reduction of predator communities with the promotion of desired bird communities and other conservation goals.}, number={2-4}, journal={Landscape and Urban Planning}, publisher={Elsevier BV}, author={Sinclair, K and Hess, G and Moorman, C and Mason, J}, year={2005}, month={Mar}, pages={277–293} } @article{vandermast_moorman_russell_van lear_2004, title={Initial vegetation response to prescribed fire in some oak-hickory forests of the South Carolina piedmont}, volume={24}, number={3}, journal={Natural Areas Journal}, author={Vandermast, D. B. and Moorman, C. E. and Russell, K. R. and Van Lear, D. H.}, year={2004}, pages={216–222} } @article{ulyshen_hanula_horn_kilgo_moorman_2004, title={Spatial and temporal patterns of beetles associated with coarse woody debris in managed bottomland hardwood forests}, volume={199}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2004.05.046}, abstractNote={Malaise traps were used to sample beetles in artificial canopy gaps of different size (0.13 ha, 0.26 ha, and 0.50 ha) and age in a South Carolina bottomland hardwood forest. Traps were placed at the center, edge, and in the surrounding forest of each gap. Young gaps (∼1 year) had large amounts of coarse woody debris compared to the surrounding forest, while older gaps (∼6 years) had virtually none. The total abundance and diversity of wood-dwelling beetles (Buprestidae, Cerambycidae, Brentidae, Bostrichidae, and Curculionidae (Scolytinae and Platypodinae)) was higher in the center of young gaps than in the center of old gaps. The abundance was higher in the center of young gaps than in the surrounding forest, while the forest surrounding old gaps and the edge of old gaps had a higher abundance and diversity of wood-dwelling beetles than did the center of old gaps. There was no difference in wood-dwelling beetle abundance between gaps of different size, but diversity was lower in 0.13 ha old gaps than in 0.26 ha or 0.50 ha old gaps. We suspect that gap size has more of an effect on woodborer abundance than indicated here because malaise traps sample a limited area. The predaceous beetle family Cleridae showed a very similar trend to that of the woodborers. Coarse woody debris is an important resource for many organisms, and our results lend further support to forest management practices that preserve coarse woody debris created during timber removal.}, number={2-3}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Ulyshen, MD and Hanula, JL and Horn, S and Kilgo, JC and Moorman, CE}, year={2004}, month={Oct}, pages={259–272} } @article{kilgo_moorman_2003, title={Patterns of cowbird parasitism in the southern Atlantic coastal plain and piedmont}, volume={115}, ISSN={["0043-5643"]}, DOI={10.1676/03-037}, abstractNote={Abstract Until recently, little information was available on patterns of brood parasitism by Brown-headed Cowbirds (Molothrus ater) in the southeastern United States, a region into which cowbirds expanded their range only during the last half of the Twentieth Century and where their abundance is relatively low. We compiled parasitism data from several published and unpublished studies conducted in Georgia and South Carolina from 1993–2000 to examine levels of brood parasitism and determine frequent host species. The combined dataset included 1,372 nests of 24 species reported in the literature to have been parasitized by cowbirds. The parasitism rate on all species combined was 8.2%. Considering only those species that served as hosts in these studies (n = 12), the parasitism rate was 9.3%. Seven species were parasitized at rates ≥10%. Based on the extent of parasitism (among studies and locations), their relative abundance, and the sample size of nests, Prairie Warblers (Dendroica discolor), Hooded Warblers (Wilsonia citrina), Yellow-breasted Chats (Icteria virens), and Indigo Buntings (Passerina cyanea), all shrub nesters, appear to be the most important cowbird hosts in the region. Parasitism on some species reported as frequent hosts elsewhere was extremely low or not documented. We conclude that the impact of brood parasitism on the seasonal fecundity of hosts in the region probably is minimal, but additional work is warranted on species of concern, such as the Painted Bunting (Passerina ciris).}, number={3}, journal={WILSON BULLETIN}, author={Kilgo, JC and Moorman, CE}, year={2003}, month={Sep}, pages={277–284} } @article{moorman_guynn_kilgo_2002, title={Hooded warbler nesting success adjacent to group-selection and clearcut edges in a southeastern bottomland forest}, volume={104}, ISSN={["1938-5129"]}, DOI={10.1650/0010-5422(2002)104[0366:HWNSAT]2.0.CO;2}, abstractNote={Abstract During the 1996, 1997, and 1998 breeding seasons, we located and monitored Hooded Warbler (Wilsonia citrina) nests in a bottomland forest and examined the effects of edge proximity, edge type, and nest-site vegetation on nesting success. Successful Hooded Warbler nests were more concealed from below and were located in nest patches with a greater abundance of >0.5-m-tall switchcane (Arundinaria gigantea) stems than unsuccessful nests. Daily nest survival rates, clutch size, and number of fledglings per successful nest did not differ between nests near edges of selection-harvest openings within the bottomland and nests near edges of clearcuts adjacent to the bottomland. Daily survival rate, clutch size, and number of fledglings per successful nest did not differ among nests 0–50 m, 51–100 m, and >100 m from the nearest edge, and probability of nest survival was not related to proximity to either edge type. However, probability of parasitism by Brown-headed Cowbirds (Molothrus ater) was higher near clearcut edges, and parasitism reduced clutch size and numbers of fledglings per successful nest. The combined effects of edge, especially edge created by the relatively small (≤0.5 ha) group-selection openings, on Hooded Warbler nesting success were minimal. However, our study was conducted in a primarily forested landscape, so cowbird abundance or negative edge effects may have been low relative to agricultural landscapes in the South. Éxito de Anidación de Wilsonia citrina en Sitios Adyacentes a Bordes de Claros Formados por Extracción de Árboles Seleccionados y por Tala Rasa en Bosques Ribereños del Sureste Resumen. Durante las épocas reproductivas de 1996, 1997 y 1998, ubicamos y monitoreamos nidos de Wilsonia citrina en un bosque de ribereño y evaluamos los efectos de la proximidad al borde, el tipo de borde y la vegetación del sitio de anidación sobre el éxito reproductivo. Los nidos exitosos estuvieron más escondidos desde abajo y se ubicaron en parches de bosque con una mayor abundancia de tallos de Arundinaria gigantea de más de 0.5 m de alto que los nidos no exitosos. Las tasas de supervivencia diaria de los nidos, el tamaño de la nidada y el número de polluelos emplumados por nido exitoso no difirieron entre nidos ubicados cerca de bordes de aperturas de cosecha selectiva dentro del valle ribereño y nidos cerca de bordes de sitios completamente talados adyacentes al valle. La tasa de supervivencia diaria, el tamaño de la nidada y el número de polluelos emplumados por nido exitoso no difirió entre nidos ubicados a 0–50 m, 51–100 m y >100 m del borde más cercano, y la probabilidad de supervivencia de los nidos no estuvo relacionada con la proximidad a ningún tipo de borde. Sin embargo, la probabilidad de parasitismo por Molothrus ater fue mayor cerca de bordes de tala rasa, y el parasitismo redujo el tamaño de la nidada y el número de polluelos emplumados por nido exitoso. Los efectos combinados de borde sobre el éxito de anidación de W. citrina fueron mínimos, especialmente aquellos de los bordes creados por los claros relativamente pequeños (≤0.5 ha) formados tras extraer grupos de árboles seleccionados. Sin embargo, nuestro estudio fue realizado en un paisaje principalmente forestal, de modo que la abundancia de M. ater o los efectos de borde negativos pueden haber sido menores en relación a paisajes agrícolas del sur.}, number={2}, journal={CONDOR}, author={Moorman, CE and Guynn, DC and Kilgo, JC}, year={2002}, month={May}, pages={366–377} } @misc{bardon_moorman_hamilton_2002, title={Response: The urbanization of North Carolina}, volume={100}, number={7}, journal={Journal of Forestry}, author={Bardon, R. E. and Moorman, C. E. and Hamilton, R. A.}, year={2002}, pages={57–58} } @article{moorman_guynn_2001, title={Effects of group-selection opening size on breeding bird habitat use in a bottomland forest}, volume={11}, ISSN={["1051-0761"]}, DOI={10.1890/1051-0761(2001)011[1680:EOGSOS]2.0.CO;2}, abstractNote={Research on the effects of creating group-selection openings of various sizes on breeding birds habitat use in a bottomland hardwood forest of the Upper Coastal Plain of South Carolina. Creation of 0.5-ha group selection openings in southern bottomland forests should provide breeding habitat for some field-edge species in gaps and habitat for forest-interior species and canopy-dwelling forest-edge species between gaps provided that enough mature forest is made available.}, number={6}, journal={ECOLOGICAL APPLICATIONS}, author={Moorman, CE and Guynn, DC}, year={2001}, month={Dec}, pages={1680–1691} } @article{moorman_howell_chapman_1999, title={Nesting ecology of Red-shouldered and Red-tailed Hawks in Georgia}, volume={33}, number={3}, journal={Journal of Raptor Research}, author={Moorman, C. E. and Howell, D. L. and Chapman, B. R.}, year={1999}, pages={248–251} }