Journal articles on the topic 'Above- and belowground'

To see the other types of publications on this topic, follow the link: Above- and belowground.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Above- and belowground.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Bennett, Alison. "Pushing boundaries in above-belowground interactions." Functional Ecology 26, no. 2 (March 27, 2012): 305–6. http://dx.doi.org/10.1111/j.1365-2435.2011.01957.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Ramirez, Kelly S., Stefan Geisen, Elly Morriën, Basten L. Snoek, and Wim H. van der Putten. "Network Analyses Can Advance Above-Belowground Ecology." Trends in Plant Science 23, no. 9 (September 2018): 759–68. http://dx.doi.org/10.1016/j.tplants.2018.06.009.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Stone, Martin J., Harry T. Cralle, James M. Chandler, Travis D. Miller, Rodney W. Bovey, and Katherine H. Carson. "Above- and belowground interference of wheat (Triticum aestivum) by Italian ryegrass (Lolium multiflorum)." Weed Science 46, no. 4 (August 1998): 438–41. http://dx.doi.org/10.1017/s004317450009086x.

Full text
Abstract:
Greenhouse experiments in central Texas assessed the relative importance of above- and belowground interactions of semidwarf Mit wheat and Marshall ryegrass during vegetative growth. One experiment used partitions to compare the effect of no (controls), aboveground only, belowground only, and full interaction for 75 d after planting (DAP) one wheat and nine ryegrass plants in soil volumes of 90, 950, and 3,800 ml. The results with the different soil volumes were similar. Wheat growth in the aboveground interaction only did not differ from controls. However, the full or belowground only interaction of wheat with ryegrass reduced wheat height, leaf number, tillering, leaf area, percent total nonstructural carbohydrates in shoot, and dry weights of leaves, stems, and roots 45 and 75 DAP compared to controls. Wheat in full and belowground interaction only did not differ from one another in growth. A replacement series experiment of 56 d also showed that the competitive advantage of ryegrass was relatively greater in root than in shoot growth. No allelopathic response of wheat to ryegrass occurred. While the tallness of the semidwarf wheat minimized aboveground interference by ryegrass, the root growth of the thinner and more fibrous roots of ryegrass greatly enhanced its belowground competitiveness.
APA, Harvard, Vancouver, ISO, and other styles
4

Cheng, J., G. L. Wu, L. P. Zhao, Y. Li, W. Li, and J. M. Cheng. "Cumulative effects of 20-year exclusion of livestock grazing on above- and belowground biomass of typical steppe communities in arid areas of the Loess Plateau, China." Plant, Soil and Environment 57, No. 1 (January 14, 2011): 40–44. http://dx.doi.org/10.17221/153/2010-pse.

Full text
Abstract:
Overgrazing affects typical steppe community in ways similar to grasslands in other areas. Exclusion of livestock grazing is one of the main management practices used to protect grasslands. However, it is not known if long-term exclusion of livestock grazing has positive effect on above- and belowground community properties in typical steppe of the Loess Plateau. We studied the long-term (20-year) cumulative effects of exclusion of livestock grazing on above- and belowground community properties compared with that before exclusion of livestock grazing in a typical steppe of the Loess Plateau, NW China. Our results show that twenty-year exclusion of livestock grazing significantly increased above- and belowground biomass, species richness, cover and height for five different communities. Most of belowground biomass was in the 0–20 cm horizon and grazing exclusion increased biomass especially at the depth of 0–10 cm. Our study suggests that long-term exclusion of livestock grazing can greatly improve community properties of typical steppe in the Loess Plateau.  
APA, Harvard, Vancouver, ISO, and other styles
5

Ye, X. H., X. Pan, W. K. Cornwell, S. Q. Gao, M. Dong, and J. H. C. Cornelissen. "Divergence of above- and belowground C and N pool within predominant plant species along two precipitation gradients in north China." Biogeosciences Discussions 11, no. 10 (October 2, 2014): 14173–95. http://dx.doi.org/10.5194/bgd-11-14173-2014.

Full text
Abstract:
Abstract. The coupling of carbon cycle and nutrient cycle drives food web structure and biogeochemistry of an ecosystem. However, across precipitation gradients, there may be a shift in C pool and N pool from above- to belowground because of shifting plant stoichiometry and allocation. Based on previous evidence, biomass allocation to roots should increase with aridity, while leaf [N] should increase. If their effect sizes are equal, they should cancel each other out, and the above- and belowground proportions of the N would remain constant. Here, we present the first study to explicitly compare above- and belowground pool sizes of N and C within predominant plant species along precipitation gradients. Biomass and nutrient concentrations of leaves, stems and roots of three predominant species were measured along two major precipitation gradients in Inner Mongolia, China. Along the two gradients, the effect sizes of the biomass shifts were remarkably consistent among three predominant species. However, the size of the shift in aboveground [N] was not, leading to a species-specific pattern in above- and belowground pool size. In two species (Stipa grandis and Artemisia ordosica) the effect sizes of biomass allocation and [N] were equal and the proportion of N of above- and belowground did not change with aridity, but in S. bungeana the increase in leaf [N] with aridity was much weaker than the biomass shift, leading to a decrease in the proportion of N belowground at dry sites. We have found examples of consistent N pool sizes above- and belowground and a shift to a greater proportion of belowground N in drier sites depending on the species. We suggest that precipitation gradients do potentially decouple the C and N pool, but the exact nature of the decoupling depends on the dominant species' capacity for intraspecific variation.
APA, Harvard, Vancouver, ISO, and other styles
6

Yang, Yuanhe, Jingyun Fang, Chengjun Ji, and Wenxuan Han. "Above- and belowground biomass allocation in Tibetan grasslands." Journal of Vegetation Science 20, no. 1 (February 2009): 177–84. http://dx.doi.org/10.1111/j.1654-1103.2009.05566.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Lyons, Caitlyn L., and Zoë Lindo. "Above- and belowground community linkages in boreal peatlands." Plant Ecology 221, no. 7 (May 20, 2020): 615–32. http://dx.doi.org/10.1007/s11258-020-01037-w.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Ribeiro, Sabina Cerruto, Lutz Fehrmann, Carlos Pedro Boechat Soares, Laércio Antônio Gonçalves Jacovine, Christoph Kleinn, and Ricardo de Oliveira Gaspar. "Above- and belowground biomass in a Brazilian Cerrado." Forest Ecology and Management 262, no. 3 (August 2011): 491–99. http://dx.doi.org/10.1016/j.foreco.2011.04.017.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Wurst, Susanne. "Effects of earthworms on above- and belowground herbivores." Applied Soil Ecology 45, no. 3 (July 2010): 123–30. http://dx.doi.org/10.1016/j.apsoil.2010.04.005.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Wang, Jin-Wang, Dan Yu, Wen Xiong, and Yu-Qin Han. "Above- and belowground competition between two submersed macrophytes." Hydrobiologia 607, no. 1 (March 28, 2008): 113–22. http://dx.doi.org/10.1007/s10750-008-9371-7.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Hagedorn, Frank, Konstantin Gavazov, and Jake M. Alexander. "Above- and belowground linkages shape responses of mountain vegetation to climate change." Science 365, no. 6458 (September 12, 2019): 1119–23. http://dx.doi.org/10.1126/science.aax4737.

Full text
Abstract:
Upward shifts of mountain vegetation lag behind rates of climate warming, partly related to interconnected changes belowground. Here, we unravel above- and belowground linkages by drawing insights from short-term experimental manipulations and elevation gradient studies. Soils will likely gain carbon in early successional ecosystems, while losing carbon as forest expands upward, and the slow, high-elevation soil development will constrain warming-induced vegetation shifts. Current approaches fail to predict the pace of these changes and how much they will be modified by interactions among plants and soil biota. Integrating mountain soils and their biota into monitoring programs, combined with innovative comparative and experimental approaches, will be crucial to overcome the paucity of belowground data and to better understand mountain ecosystem dynamics and their feedbacks to climate.
APA, Harvard, Vancouver, ISO, and other styles
12

Ye, X. H., X. Pan, W. K. Cornwell, S. Q. Gao, M. Dong, and J. H. C. Cornelissen. "Divergence of above- and belowground C and N pool within predominant plant species along two precipitation gradients in North China." Biogeosciences 12, no. 2 (January 27, 2015): 457–65. http://dx.doi.org/10.5194/bg-12-457-2015.

Full text
Abstract:
Abstract. The coupling of carbon cycle and nitrogen cycle drives the food web structure and biogeochemistry of an ecosystem. However, across precipitation gradients, there may be a shift in C pool and N pool from above- to belowground because of shifting plant stoichiometry and allocation. Based on previous evidence, biomass allocation to roots should increase with aridity, while leaf [N] should increase. If their effect sizes are equal, they should cancel each other out, and the above- and belowground proportions of the N would remain constant. Here, we present the first study to explicitly compare above- and belowground pool sizes of N and C within predominant plant species along precipitation gradients. Biomass and nutrient concentrations of leaves, stems and roots of three predominant species were measured along two major precipitation gradients in Inner Mongolia, China. Along the two gradients, the effect sizes of the biomass shifts were remarkably consistent among three predominant species. However, the size of the shift in aboveground [N] was not, leading to a species-specific pattern in above- and belowground pool size. In two species (Stipa grandis and Artemisia ordosica) the effect sizes of biomass allocation and [N] were equal and the proportion of N of above- and belowground did not change with aridity, but in S. bungeana the increase in leaf [N] with aridity was much weaker than the biomass shift, leading to a decrease in the proportion of N aboveground at dry sites. We have found examples of consistent N pool sizes above- and belowground and a shift to a greater proportion of belowground N in drier sites depending on the species. We suggest that precipitation gradients do potentially decouple the C and N pool, but the exact nature of the decoupling depends on the dominant species' capacity for intraspecific variation.
APA, Harvard, Vancouver, ISO, and other styles
13

Haase, Josephine, Roland Brandl, Stefan Scheu, and Martin Schädler. "ABOVE‐ AND BELOWGROUND INTERACTIONS ARE MEDIATED BY NUTRIENT AVAILABILITY." Ecology 89, no. 11 (November 2008): 3072–81. http://dx.doi.org/10.1890/07-1983.1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Jamieson, Mary A., Timothy R. Seastedt, and M. Deane Bowers. "Nitrogen enrichment differentially affects above- and belowground plant defense." American Journal of Botany 99, no. 10 (October 2012): 1630–37. http://dx.doi.org/10.3732/ajb.1100492.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Roeder, Karl A., Diane V. Roeder, and Michael Kaspari. "Disturbance Mediates Homogenization of Above and Belowground Invertebrate Communities." Environmental Entomology 47, no. 3 (March 15, 2018): 545–50. http://dx.doi.org/10.1093/ee/nvy022.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Deyn, Gerlinde B. De. "Plant life history and above-belowground interactions: missing links." Oikos 126, no. 4 (February 1, 2017): 497–507. http://dx.doi.org/10.1111/oik.03967.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Roy, Austin, Matthew Suchocki, Laura Gough, and Jennie R. McLaren. "Above- and belowground responses to long-term herbivore exclusion." Arctic, Antarctic, and Alpine Research 52, no. 1 (January 1, 2020): 109–19. http://dx.doi.org/10.1080/15230430.2020.1733891.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Tan, Xinyuan, Hong He, Shengwei Zong, Miaomiao Wu, Kai Liu, and Dandan Zhao. "Herbaceous Encroachment from Mountain Birch Forests to Alpine Tundra Plant Communities Through Above- and Belowground Competition." Forests 10, no. 2 (February 16, 2019): 170. http://dx.doi.org/10.3390/f10020170.

Full text
Abstract:
Alpine plant communities are highly sensitive to global warming. One of the consequences of the warming is encroachment by herbaceous plants from forests at low elevations into alpine ecosystems. In the Changbai Mountains, narrowleaf small reed (Deyeuxia angustifolia (Kom.) Y. L. Chang) from mountain birch forests encroached upward into alpine tundra, gradually replacing native tundra shrubs such as Rhododendron (Rhododendron aureum Georgi). How encroaching plants affect native plant communities is not fully understood. In this study, we analyzed above- and belowground biomass of alpine plant communities at five encroachment levels to investigate how biomass allocation changed at species and community scales. Our research showed that native plants are forced to change their morphology to cope with competition, at both above- and belowground levels, from encroaching plants. We found that (1) R. aureum increased the shoot height and leaf area in order to compete with D. angustifolia; (2) above- and belowground biomass of D. angustifolia increased while above- and belowground biomass of R. aureum decreased with increasing levels of encroachment; and (3) D. angustifolia encroachment reduced the total biomass of alpine tundra. Encroachment by herbaceous plants has a long-term negative impact on the ability of tundra plants to sequester carbon in the alpine tundra of the Changbai Mountains.
APA, Harvard, Vancouver, ISO, and other styles
19

Ye, X. H., X. Pan, W. K. Cornwell, J. H. C. Cornelissen, Y. Chu, S. Q. Gao, R. Q. Li, J. J. Qiao, and M. Dong. "Decoupling of above and belowground C and N pools within predominant plant species <i>Stipa grandis</i> along a precipitation gradient in Chinese steppe zone." Biogeosciences Discussions 10, no. 3 (March 12, 2013): 4995–5013. http://dx.doi.org/10.5194/bgd-10-4995-2013.

Full text
Abstract:
Abstract. The coupling of the carbon and nutrient cycles drives the food web structure and biogeochemistry of ecosystems. However, across precipitation gradients, there may be a shift in C and N pools from above- to belowground because of shifting plant stoichiometry and allocation. Here, we present a study which is the first to explicitly compare above- and belowground pool sizes of N and C within predominant plant species along precipitation gradient. We dissected these pools into biomass allocation and nutrient concentrations. Based on previous evidence, biomass allocation to roots should increase with aridity, while leaf [N] should increase. If their effect sizes are equal, they should cancel each other out, and the above- and belowground proportions of the N would remain constant. Along a precipitation gradient in Chinese steppe zone, the effect sizes of the biomass shifts were remarkably consistent among the predominant species, Stipa grandis. The effect sizes of biomass allocation and [N] were equal and the proportion of N of above- and belowground did not change with aridity, but the shift in leaf [C] with aridity was much weaker than the biomass shift, leading to a decrease in the proportion of C belowground at dry sites. Precipitation gradients do decouple the C and N pool of S. grandis along a precipitation gradient in Chinese steppe zone.
APA, Harvard, Vancouver, ISO, and other styles
20

Zhao, Li, Zhou, Qiu, and Wu. "Site-Specific Allometric Models for Prediction of Above-and Belowground Biomass of Subtropical Forests in Guangzhou, Southern China." Forests 10, no. 10 (October 2, 2019): 862. http://dx.doi.org/10.3390/f10100862.

Full text
Abstract:
Tree allometric models that are used to predict the biomass of individual tree are critical to forest carbon accounting and ecosystem service modeling. To enhance the accuracy of such predictions, the development of site-specific, rather than generalized, allometric models is advised whenever possible. Subtropical forests are important carbon sinks and have a huge potential for mitigating climate change. However, few biomass models compared to the diversity of forest ecosystems are currently available for the subtropical forests of China. This study developed site-specific allometric models to estimate the aboveground and the belowground biomass for south subtropical humid forest in Guangzhou, Southern China. Destructive methods were used to measure the aboveground biomass with a sample of 144 trees from 26 species, and the belowground biomass was measured with a subsample of 116 of them. Linear regression with logarithmic transformation was used to model biomass according to dendrometric parameters. The mixed-species regressions with diameter at breast height (DBH) as a single predictor were able to adequately estimate aboveground, belowground and total biomass. The coefficients of determination (R2) were 0.955, 0.914 and 0.954, respectively, and the mean prediction errors were −1.96, −5.84 and 2.26%, respectively. Adding tree height (H) compounded with DBH as one variable (DBH2H) did not improve model performance. Using H as a second variable in the equation can improve the model fitness in estimation of belowground biomass, but there are collinearity effects, resulting in an increased standard error of regression coefficients. Therefore, it is not recommended to add H in the allometric models. Adding wood density (WD) compounded with DBH as one variable (DBH2WD) slightly improved model fitness for prediction of belowground biomass, but there was no positive effect on the prediction of aboveground and total biomass. Using WD as a second variable in the equation, the best-fitting allometric relationship for biomass estimation of the aboveground, belowground, and total biomass was given, indicating that WD is a crucial factor in biomass models of subtropical forest. Root-shoot ratio of subtropical forest in this study varies with species and tree size, and it is not suitable to apply it to estimate belowground biomass. These findings are of great significance for accurately measuring regional forest carbon sinks, and having reference value for forest management.
APA, Harvard, Vancouver, ISO, and other styles
21

Jargalsaikhan, Gantuya. "A review of similarity between seed bank and standing vegetation under grazing." Mongolian Journal of Agricultural Sciences 11, no. 2 (November 25, 2014): 191–96. http://dx.doi.org/10.5564/mjas.v11i2.243.

Full text
Abstract:
In recent years, many researchers have stated the importance of above and belowground interactions to better understand succession in plant communities and state and transition dynamics in rangelands. A review indicate that improved knowledge the soil's seed bank is a key element in understanding above and belowground interactions and plant community dynamics in grazed rangelands. The aim was to study current successional theories, with special emphasis on state and transition models to understand rangeland ecosystem dynamics under grazing. I thoroughly reviewed 28 articles published that summarized and provided specific values on similarities between above and belowground communities to identify under grazing across different ecosystem DOI: http://dx.doi.org/10.5564/mjas.v11i2.243 Mongolian Journal of Agricultural Sciences Vol.11(2) 2013 pp.191-196
APA, Harvard, Vancouver, ISO, and other styles
22

Wangchuk, Kesang, Andras Darabant, and Prem Bahadur Rai. "Morphological Responses Explain Tolerance of the Bamboo Yushania microphylla to Grazing." Journal of Botany 2014 (August 19, 2014): 1–7. http://dx.doi.org/10.1155/2014/573415.

Full text
Abstract:
Mechanisms of tolerance of the bamboo Y. microphylla to ungulate herbivory were investigated by measuring above- and belowground morphogenetic traits and biomass allocation patterns of the bamboo Y. microphylla under grazed and ungrazed conditions in a Himalayan mixed conifer forest. Data were collected from 5 populations consisting of 10 ramets each in adjacent grazed and ungrazed plots. Compared with ungrazed ramets, the aboveground morphological modifications of grazed ramets were higher culm density, shorter and thinner culms, shorter internode, and shorter top leaf. The belowground morphological modifications for the grazed ramets were thinner rhizomes, lower rhizome biomass and dry matter, more nodes, and shorter internodes. Despite the lower biomass and dry matter, the root-to-shoot ratio was higher for grazed ramets. Results suggest that Y. microphylla subjected to herbivory shows aboveground overcompensation in terms of densification at the cost of belowground biomass, but at the same time maintains a higher proportion of belowground reserves, as compared to ungrazed conditions. These responses provide adequate evidence to conclude that Y. microphylla tolerates ungulate herbivory through above- and belowground morphological modifications.
APA, Harvard, Vancouver, ISO, and other styles
23

Marella, Venkata S. S. R., Paul W. Hill, Davey L. Jones, and Paula Roberts. "Microbial turnover of above and belowground litter components in shrublands." Pedobiologia 59, no. 4 (July 2016): 229–32. http://dx.doi.org/10.1016/j.pedobi.2016.07.001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Pornaro, C., M. K. Schneider, B. Leinauer, and S. Macolino. "Above- and belowground patterns in a subalpine grassland-shrub mosaic." Plant Biosystems - An International Journal Dealing with all Aspects of Plant Biology 151, no. 3 (June 2, 2016): 493–503. http://dx.doi.org/10.1080/11263504.2016.1187679.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

LEAKE, J. R., and D. D. CAMERON. "Untangling above- and belowground mycorrhizal fungal networks in tropical orchids." Molecular Ecology 21, no. 20 (October 2012): 4921–24. http://dx.doi.org/10.1111/j.1365-294x.2012.05718.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Delgado-Baquerizo, Manuel, David J. Eldridge, Samantha K. Travers, James Val, Ian Oliver, and Andrew Bissett. "Effects of climate legacies on above- and belowground community assembly." Global Change Biology 24, no. 9 (May 30, 2018): 4330–39. http://dx.doi.org/10.1111/gcb.14306.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Liu, Weiguo, Xiuhua Fan, Jinsong Wang, Chunyu Zhang, Wenmin Lu, and Klaus V. Gadow. "Spectral reflectance response ofFraxinus mandshuricaleaves to above- and belowground competition." International Journal of Remote Sensing 33, no. 16 (February 16, 2012): 5072–86. http://dx.doi.org/10.1080/01431161.2012.657371.

Full text
APA, Harvard, Vancouver, ISO, and other styles
28

Mao, Mao, Steven A. Cryer, Anthony Altieri, and Patrick Havens. "Predicting Pesticide Volatility Through Coupled Above- and Belowground Multiphysics Modeling." Environmental Modeling & Assessment 23, no. 5 (March 7, 2018): 569–82. http://dx.doi.org/10.1007/s10666-018-9594-6.

Full text
APA, Harvard, Vancouver, ISO, and other styles
29

Visioli, Giovanna, Anna Maria Sanangelantoni, Federica D. Conti, Beatrice Bonati, Ciro Gardi, and Cristina Menta. "Above and belowground biodiversity in adjacent and distinct serpentine soils." Applied Soil Ecology 133 (January 2019): 98–103. http://dx.doi.org/10.1016/j.apsoil.2018.09.013.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

Li, Yang, Shiyu Zhen, Shaojie Shan, Bingjiao Sun, Jingjing Li, Fangzhong Hu, Qingxin Cui, et al. "Modulation of above-belowground plant-herbivore interactions by entomopathogenic nematodes." Applied Soil Ecology 148 (April 2020): 103479. http://dx.doi.org/10.1016/j.apsoil.2019.103479.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Menacer, Kathleen, Anne Marie Cortesero, and Maxime R. Hervé. "Challenging the Preference–Performance Hypothesis in an above-belowground insect." Oecologia 197, no. 1 (August 7, 2021): 179–87. http://dx.doi.org/10.1007/s00442-021-05007-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Smith, Jennifer E., Denisse A. Gamboa, Julia M. Spencer, Sarah J. Travenick, Chelsea A. Ortiz, Riana D. Hunter, and Andy Sih. "Split between two worlds: automated sensing reveals links between above- and belowground social networks in a free-living mammal." Philosophical Transactions of the Royal Society B: Biological Sciences 373, no. 1753 (July 2, 2018): 20170249. http://dx.doi.org/10.1098/rstb.2017.0249.

Full text
Abstract:
Many animals socialize in two or more major ecological contexts. In nature, these contexts often involve one situation in which space is more constrained (e.g. shared refuges, sleeping cliffs, nests, dens or burrows) and another situation in which animal movements are relatively free (e.g. in open spaces lacking architectural constraints). Although it is widely recognized that an individual's characteristics may shape its social life, the extent to which architecture constrains social decisions within and between habitats remains poorly understood. Here we developed a novel, automated-monitoring system to study the effects of personality, life-history stage and sex on the social network structure of a facultatively social mammal, the California ground squirrel ( Otospermophilus beecheyi ) in two distinct contexts: aboveground where space is relatively open and belowground where it is relatively constrained by burrow architecture. Aboveground networks reflected affiliative social interactions whereas belowground networks reflected burrow associations. Network structure in one context (belowground), along with preferential juvenile–adult associations, predicted structure in a second context (aboveground). Network positions of individuals were generally consistent across years (within contexts) and between ecological contexts (within years), suggesting that individual personalities and behavioural syndromes, respectively, contribute to the social network structure of these free-living mammals. Direct ties (strength) tended to be stronger in belowground networks whereas more indirect paths (betweenness centrality) flowed through individuals in aboveground networks. Belowground, females fostered significantly more indirect paths than did males. Our findings have important potential implications for disease and information transmission, offering new insights into the multiple factors contributing to social structures across ecological contexts. This article is part of the theme issue ‘Interdisciplinary approaches for uncovering the impacts of architecture on collective behaviour’.
APA, Harvard, Vancouver, ISO, and other styles
33

Dirks, Inga, Juliane Streit, and Catharina Meinen. "Above and Belowground Relative Yield Total of Clover–Ryegrass Mixtures Exceed One in Wet and Dry Years." Agriculture 11, no. 3 (March 3, 2021): 206. http://dx.doi.org/10.3390/agriculture11030206.

Full text
Abstract:
Grassland mixtures hold the potential for increasing biomass and productivity. In a field experiment, monocultures and mixtures of eight white clover (Trifolium repens L.) genotypes and perennial ryegrass (Lolium perenne L.) were analyzed over three years (2015, 2016, and 2018) for their species-specific aboveground and belowground biomass. Roots were analyzed by Fourier transform infrared (FTIR) spectroscopy to identify species-specific root mass, vertical distribution, and belowground relative yield total (RYT). Aboveground biomass decreased strongly from 2015 to 2018. Aboveground and belowground RYT were always significantly higher than one. Aboveground biomass overyielded in 2016 and 2018 compared to monocultures. Monocultures of perennial ryegrass displayed a significantly higher proportion of roots in shallow soil layers than white clover in two of the three examined years. In mixtures, these differences in vertical root distribution between both species were not present and perennial ryegrass, and white clover occupied similar vertical niches in 2015 and 2016. Interestingly, in the dry year 2018, white clover had a higher proportion of roots in shallow soil layers than perennial ryegrass in mixtures.
APA, Harvard, Vancouver, ISO, and other styles
34

Korell, Lotte, Martin Schädler, Roland Brandl, Susanne Schreiter, and Harald Auge. "Release from Above- and Belowground Insect Herbivory Mediates Invasion Dynamics and Impact of an Exotic Plant." Plants 8, no. 12 (November 26, 2019): 544. http://dx.doi.org/10.3390/plants8120544.

Full text
Abstract:
The enemy-release hypothesis is one of the most popular but also most discussed hypotheses to explain invasion success. However, there is a lack of explicit, experimental tests of predictions of the enemy-release hypothesis (ERH), particularly regarding the effects of above- and belowground herbivory. Long-term studies investigating the relative effect of herbivores on invasive vs. native plant species within a community are still lacking. Here, we report on a long-term field experiment in an old-field community, invaded by Solidago canadensis s. l., with exclusion of above- and belowground insect herbivores. We monitored population dynamics of the invader and changes in the diversity and functioning of the plant community across eight years. Above- and belowground insects favoured the establishment of the invasive plant species and thereby increased biomass and decreased diversity of the plant community. Effects of invertebrate herbivores on population dynamics of S. canadensis appeared after six years and increased over time, suggesting that long-term studies are needed to understand invasion dynamics and consequences for plant community structure. We suggest that the release from co-evolved trophic linkages is of importance not only for the effect of invasive species on ecosystems, but also for the functioning of novel species assemblages arising from climate change.
APA, Harvard, Vancouver, ISO, and other styles
35

Webb, Elizabeth E., Kathryn Heard, Susan M. Natali, Andrew G. Bunn, Heather D. Alexander, Logan T. Berner, Alexander Kholodov, et al. "Variability in above- and belowground carbon stocks in a Siberian larch watershed." Biogeosciences 14, no. 18 (September 26, 2017): 4279–94. http://dx.doi.org/10.5194/bg-14-4279-2017.

Full text
Abstract:
Abstract. Permafrost soils store between 1330 and 1580 Pg carbon (C), which is 3 times the amount of C in global vegetation, almost twice the amount of C in the atmosphere, and half of the global soil organic C pool. Despite the massive amount of C in permafrost, estimates of soil C storage in the high-latitude permafrost region are highly uncertain, primarily due to undersampling at all spatial scales; circumpolar soil C estimates lack sufficient continental spatial diversity, regional intensity, and replication at the field-site level. Siberian forests are particularly undersampled, yet the larch forests that dominate this region may store more than twice as much soil C as all other boreal forest types in the continuous permafrost zone combined. Here we present above- and belowground C stocks from 20 sites representing a gradient of stand age and structure in a larch watershed of the Kolyma River, near Chersky, Sakha Republic, Russia. We found that the majority of C stored in the top 1 m of the watershed was stored belowground (92 %), with 19 % in the top 10 cm of soil and 40 % in the top 30 cm. Carbon was more variable in surface soils (10 cm; coefficient of variation (CV) = 0.35 between stands) than in the top 30 cm (CV = 0.14) or soil profile to 1 m (CV = 0.20). Combined active-layer and deep frozen deposits (surface – 15 m) contained 205 kg C m−2 (yedoma, non-ice wedge) and 331 kg C m−2 (alas), which, even when accounting for landscape-level ice content, is an order of magnitude more C than that stored in the top meter of soil and 2 orders of magnitude more C than in aboveground biomass. Aboveground biomass was composed of primarily larch (53 %) but also included understory vegetation (30 %), woody debris (11 %) and snag (6 %) biomass. While aboveground biomass contained relatively little (8 %) of the C stocks in the watershed, aboveground processes were linked to thaw depth and belowground C storage. Thaw depth was negatively related to stand age, and soil C density (top 10 cm) was positively related to soil moisture and negatively related to moss and lichen cover. These results suggest that, as the climate warms, changes in stand age and structure may be as important as direct climate effects on belowground environmental conditions and permafrost C vulnerability.
APA, Harvard, Vancouver, ISO, and other styles
36

Armitage, A. R., and J. W. Fourqurean. "Carbon storage in seagrass soils: long-term nutrient history exceeds the effects of near-term nutrient enrichment." Biogeosciences 13, no. 1 (January 15, 2016): 313–21. http://dx.doi.org/10.5194/bg-13-313-2016.

Full text
Abstract:
Abstract. The carbon sequestration potential in coastal soils is linked to aboveground and belowground plant productivity and biomass, which in turn, is directly and indirectly influenced by nutrient input. We evaluated the influence of long-term and near-term nutrient input on aboveground and belowground carbon accumulation in seagrass beds, using a nutrient enrichment (nitrogen and phosphorus) experiment embedded within a naturally occurring, long-term gradient of phosphorus availability within Florida Bay (USA). We measured organic carbon stocks in soils and above- and belowground seagrass biomass after 17 months of experimental nutrient addition. At the nutrient-limited sites, phosphorus addition increased the carbon stock in aboveground seagrass biomass by more than 300 %; belowground seagrass carbon stock increased by 50–100 %. Soil carbon content slightly decreased ( ∼ 10 %) in response to phosphorus addition. There was a strong but non-linear relationship between soil carbon and Thalassia testudinum leaf nitrogen : phosphorus (N : P) or belowground seagrass carbon stock. When seagrass leaf N : P exceeded an approximate threshold of 75 : 1, or when belowground seagrass carbon stock was less than 100 g m−2, there was less than 3 % organic carbon in the sediment. Despite the marked difference in soil carbon between phosphorus-limited and phosphorus-replete areas of Florida Bay, all areas of the bay had relatively high soil carbon stocks near or above the global median of 1.8 % organic carbon. The relatively high carbon content in the soils indicates that seagrass beds have extremely high carbon storage potential, even in nutrient-limited areas with low biomass or productivity.
APA, Harvard, Vancouver, ISO, and other styles
37

Armitage, A. R., and J. W. Fourqurean. "Carbon storage in seagrass soils: long-term nutrient history exceeds the effects of near-term nutrient enrichment." Biogeosciences Discussions 12, no. 19 (October 2, 2015): 16285–312. http://dx.doi.org/10.5194/bgd-12-16285-2015.

Full text
Abstract:
Abstract. The carbon sequestration potential in coastal soils is linked to aboveground and belowground plant productivity and biomass, which in turn, is directly and indirectly influenced by nutrient input. We evaluated the influence of long-term and near-term nutrient input on aboveground and belowground carbon accumulation in seagrass beds, using a nutrient enrichment (nitrogen and phosphorus) experiment embedded within a naturally occurring, long-term gradient of phosphorus availability within Florida Bay (USA). We measured organic carbon stocks in soils and above- and belowground seagrass biomass after 17 months of experimental nutrient addition. At the nutrient-limited sites, phosphorus addition increased the carbon stock in aboveground seagrass biomass by more than 300 %; belowground seagrass carbon stock increased by 50–100 %. Soil carbon content slightly decreased (~ 10 %) in response to phosphorus addition. There was a strong but non-linear relationship between soil carbon and Thalassia testudinum leaf nitrogen: phosphorus (N : P) or belowground seagrass carbon stock. When seagrass leaf N : P exceeded a threshold of 75 : 1, or when belowground seagrass carbon stock was less than 100 g m−2, there was less than 3 % organic carbon in the sediment. Despite the marked difference in soil carbon between phosphorus-limited and phosphorus-replete areas of Florida Bay, all areas of the bay had relatively high soil carbon stocks near or above the global median of 1.8 % organic carbon. The relatively high carbon content in the soils indicates that seagrass beds have extremely high carbon storage potential, even in nutrient-limited areas with low biomass or productivity.
APA, Harvard, Vancouver, ISO, and other styles
38

Sierra Cornejo, Natalia, Christoph Leuschner, Joscha N. Becker, Andreas Hemp, David Schellenberger Costa, and Dietrich Hertel. "Climate implications on forest above- and belowground carbon allocation patterns along a tropical elevation gradient on Mt. Kilimanjaro (Tanzania)." Oecologia 195, no. 3 (February 25, 2021): 797–812. http://dx.doi.org/10.1007/s00442-021-04860-8.

Full text
Abstract:
AbstractTropical forests represent the largest store of terrestrial biomass carbon (C) on earth and contribute over-proportionally to global terrestrial net primary productivity (NPP). How climate change is affecting NPP and C allocation to tree components in forests is not well understood. This is true for tropical forests, but particularly for African tropical forests. Studying forest ecosystems along elevation and related temperature and moisture gradients is one possible approach to address this question. However, the inclusion of belowground productivity data in such studies is scarce. On Mt. Kilimanjaro (Tanzania), we studied aboveground (wood increment, litter fall) and belowground (fine and coarse root) NPP along three elevation transects (c. 1800–3900 m a.s.l.) across four tropical montane forest types to derive C allocation to the major tree components. Total NPP declined continuously with elevation from 8.5 to 2.8 Mg C ha−1 year−1 due to significant decline in aboveground NPP, while fine root productivity (sequential coring approach) remained unvaried with around 2 Mg C ha−1 year−1, indicating a marked shift in C allocation to belowground components with elevation. The C and N fluxes to the soil via root litter were far more important than leaf litter inputs in the subalpine Erica forest. Thus, the shift of C allocation to belowground organs with elevation at Mt. Kilimanjaro and other tropical forests suggests increasing nitrogen limitation of aboveground tree growth at higher elevations. Our results show that studying fine root productivity is crucial to understand climate effects on the carbon cycle in tropical forests.
APA, Harvard, Vancouver, ISO, and other styles
39

Twolan-Strutt, Lisa, and Paul A. Keddy. "Above- and Belowground Competition Intensity in Two Contrasting Wetland Plant Communities." Ecology 77, no. 1 (January 1996): 259–70. http://dx.doi.org/10.2307/2265675.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Zhang, Jun, and James T. Romo. "Defoliation of a Northern Wheatgrass Community: Above- and Belowground Phytomass Productivity." Journal of Range Management 47, no. 4 (July 1994): 279. http://dx.doi.org/10.2307/4002548.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Stevnbak, Karen, Christoph Scherber, David J. Gladbach, Claus Beier, Teis N. Mikkelsen, and Søren Christensen. "Interactions between above- and belowground organisms modified in climate change experiments." Nature Climate Change 2, no. 11 (May 20, 2012): 805–8. http://dx.doi.org/10.1038/nclimate1544.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Carvalho, Joao L. N., Tara W. Hudiburg, Henrique C. J. Franco, and Evan H. DeLucia. "Contribution of above- and belowground bioenergy crop residues to soil carbon." GCB Bioenergy 9, no. 8 (January 7, 2017): 1333–43. http://dx.doi.org/10.1111/gcbb.12411.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Huynh, Trinh, David J. Lee, Grahame Applegate, and Tom Lewis. "Field methods for above and belowground biomass estimation in plantation forests." MethodsX 8 (2021): 101192. http://dx.doi.org/10.1016/j.mex.2020.101192.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Skinner, R. Howard, Matt A. Sanderson, Benjamin F. Tracy, and Curtis J. Dell. "Above- and Belowground Productivity and Soil Carbon Dynamics of Pasture Mixtures." Agronomy Journal 98, no. 2 (March 2006): 320–26. http://dx.doi.org/10.2134/agronj2005.0180a.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Vu, Huy D., and Steven C. Pennings. "Predators mediate above- vs. belowground herbivory in a salt marsh crab." Ecosphere 9, no. 2 (February 2018): e02107. http://dx.doi.org/10.1002/ecs2.2107.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Vandegehuchte, Martijn L., Eduardo de la Peña, and Dries Bonte. "Contrasting covariation of above- and belowground invertebrate species across plant genotypes." Journal of Animal Ecology 80, no. 1 (October 21, 2010): 148–58. http://dx.doi.org/10.1111/j.1365-2656.2010.01766.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Pedersen, Jane Kongstad, M. F. Arndal, and I. K. Schmidt. "Above and belowground phenology in a heathland during future climate change." IOP Conference Series: Earth and Environmental Science 6, no. 31 (February 1, 2009): 312011. http://dx.doi.org/10.1088/1755-1307/6/31/312011.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Jassey, Vincent EJ, Geneviève Chiapusio, Philippe Binet, Alexandre Buttler, Fatima Laggoun-Défarge, Frédéric Delarue, Nadine Bernard, et al. "Above- and belowground linkages inSphagnumpeatland: climate warming affects plant-microbial interactions." Global Change Biology 19, no. 3 (December 15, 2012): 811–23. http://dx.doi.org/10.1111/gcb.12075.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Rasmann, Sergio, Alison Bennett, Arjen Biere, Alison Karley, and Emilio Guerrieri. "Root symbionts: Powerful drivers of plant above- and belowground indirect defenses." Insect Science 24, no. 6 (July 3, 2017): 947–60. http://dx.doi.org/10.1111/1744-7917.12464.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Slabbert, Eleonore L., Oliver Schweiger, Tesfaye Wubet, Antje Kautzner, Cornelia Baessler, Harald Auge, Christiane Roscher, and Tiffany M. Knight. "Scale‐dependent impact of land management on above‐ and belowground biodiversity." Ecology and Evolution 10, no. 18 (August 31, 2020): 10139–49. http://dx.doi.org/10.1002/ece3.6675.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography