@article{SollySchoeningBochetal.2014, author = {Solly, Emily F. and Sch{\"o}ning, Ingo and Boch, Steffen and Kandeler, Ellen and Marhan, Sven and Michalzik, Beate and M{\"u}ller, J{\"o}rg and Zscheischler, Jakob and Trumbore, Susan E. and Schrumpf, Marion}, title = {Factors controlling decomposition rates of fine root litter in temperate forests and grasslands}, series = {Plant and soil}, volume = {382}, journal = {Plant and soil}, number = {1-2}, publisher = {Springer}, address = {Dordrecht}, issn = {0032-079X}, doi = {10.1007/s11104-014-2151-4}, pages = {203 -- 218}, year = {2014}, abstract = {Fine root decomposition contributes significantly to element cycling in terrestrial ecosystems. However, studies on root decomposition rates and on the factors that potentially influence them are fewer than those on leaf litter decomposition. To study the effects of region and land use intensity on fine root decomposition, we established a large scale study in three German regions with different climate regimes and soil properties. Methods In 150 forest and 150 grassland sites we deployed litterbags (100 mu m mesh size) with standardized litter consisting of fine roots from European beech in forests and from a lowland mesophilous hay meadow in grasslands. In the central study region, we compared decomposition rates of this standardized litter with root litter collected on-site to separate the effect of litter quality from environmental factors. Standardized herbaceous roots in grassland soils decomposed on average significantly faster (24 +/- 6 \% mass loss after 12 months, mean +/- SD) than beech roots in forest soils (12 +/- 4 \%; p < 0.001). Fine root decomposition varied among the three study regions. Land use intensity, in particular N addition, decreased fine root decomposition in grasslands. The initial lignin:N ratio explained 15 \% of the variance in grasslands and 11 \% in forests. Soil moisture, soil temperature, and C:N ratios of soils together explained 34 \% of the variance of the fine root mass loss in grasslands, and 24 \% in forests. Grasslands, which have higher fine root biomass and root turnover compared to forests, also have higher rates of root decomposition. Our results further show that at the regional scale fine root decomposition is influenced by environmental variables such as soil moisture, soil temperature and soil nutrient content. Additional variation is explained by root litter quality.}, language = {en} } @misc{Mueller1983, author = {M{\"u}ller, Richard G. E.}, title = {Klauer, Karl Josef (Hrsg.): Handbuch der P{\"a}dagogischen Diagnostik / [rezensiert von] Richard G. E. M{\"u}ller}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-59518}, year = {1983}, abstract = {rezensiertes Werk: Klauer, Karl Josef (Hrsg.): Handbuch der P{\"a}dagogischen Diagnostik. - Bd. 1-2. - D{\"u}sseldorf : P{\"a}dagogischer Verlag Schwann, 1978 Bd. 1: ISBN 3-590-14425-4 Bd. 2: ISBN 3-590-14426-2}, language = {de} } @misc{Mueller1983, author = {M{\"u}ller, Richard G. E.}, title = {Wieczerkowski, Wilhelm ; Oeveste, Hans zur: Lehrbuch der Entwicklungspsychologie / [rezensiert von] Richard G.E. M{\"u}ller}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-59529}, year = {1983}, abstract = {rezensiertes Werk: Wieczerkowski, Wilhelm ; Oeveste, Hans zur: Lehrbuch der Entwicklungspsychologie. - Bd. 1-3. - D{\"u}sseldorf : P{\"a}dagogischer Verlag Schwann, 1982 Bd. 1: ISBN 3-590-14386-X Bd. 2: ISBN 3-590-14387-8 Bd. 3: ISBN 3-590-14388-6}, language = {de} } @misc{Mueller1983, author = {M{\"u}ller, Richard G. E.}, title = {Kerkhoff, Engelbert: Handbuch Praxis der Sozialarbeit und Sozialp{\"a}dagogik / [rezensiert von] Richard G. E. M{\"u}ller}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-59535}, year = {1983}, abstract = {rezensiertes Werk: Kerkhoff, Engelbert: Handbuch Praxis der Sozialarbeit und Sozialp{\"a}dagogik. - Bd. 1-2. - D{\"u}sseldorf : P{\"a}dagogischer Verlag Schwann, 1981 Bd. 1: ISBN 3-590-14379-7 Bd. 2: ISBN 3-590-14380-0}, language = {de} } @misc{Mueller1983, author = {M{\"u}ller, Richard G. E.}, title = {Merz, Karl: Kinder mit Schulschwierigkeiten. Empirische Untersuchungen an Grund- und Sonderschulen / [rezensiert von] Richard G. E. M{\"u}ller}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-59548}, year = {1983}, abstract = {rezensiertes Werk: Merz, Karl: Kinder mit Schulschwierigkeiten. Empirische Untersuchungen an Grund- und Sonderschulen. - Weinheim u.a. : Beltz, 1982. - 395 S. ISBN 3-407-58151-3}, language = {de} } @article{SoliveresvanderPlasManningetal.2016, author = {Soliveres, Santiago and van der Plas, Fons and Manning, Peter and Prati, Daniel and Gossner, Martin M. and Renner, Swen C. and Alt, Fabian and Arndt, Hartmut and Baumgartner, Vanessa and Binkenstein, Julia and Birkhofer, Klaus and Blaser, Stefan and Bl{\"u}thgen, Nico and Boch, Steffen and B{\"o}hm, Stefan and B{\"o}rschig, Carmen and Buscot, Francois and Diek{\"o}tter, Tim and Heinze, Johannes and H{\"o}lzel, Norbert and Jung, Kirsten and Klaus, Valentin H. and Kleinebecker, Till and Klemmer, Sandra and Krauss, Jochen and Lange, Markus and Morris, E. Kathryn and M{\"u}ller, J{\"o}rg and Oelmann, Yvonne and Overmann, J{\"o}rg and Pasalic, Esther and Rillig, Matthias C. and Schaefer, H. Martin and Schloter, Michael and Schmitt, Barbara and Sch{\"o}ning, Ingo and Schrumpf, Marion and Sikorski, Johannes and Socher, Stephanie A. and Solly, Emily F. and Sonnemann, Ilja and Sorkau, Elisabeth and Steckel, Juliane and Steffan-Dewenter, Ingolf and Stempfhuber, Barbara and Tschapka, Marco and T{\"u}rke, Manfred and Venter, Paul C. and Weiner, Christiane N. and Weisser, Wolfgang W. and Werner, Michael and Westphal, Catrin and Wilcke, Wolfgang and Wolters, Volkmar and Wubet, Tesfaye and Wurst, Susanne and Fischer, Markus and Allan, Eric}, title = {Biodiversity at multiple trophic levels is needed for ecosystem multifunctionality}, series = {Nature : the international weekly journal of science}, volume = {536}, journal = {Nature : the international weekly journal of science}, publisher = {Nature Publ. Group}, address = {London}, issn = {0028-0836}, doi = {10.1038/nature19092}, pages = {456 -- +}, year = {2016}, language = {en} } @article{ZwickelKahlKlaffkeetal.2016, author = {Zwickel, Theresa and Kahl, Sandra M. and Klaffke, Horst and Rychlik, Michael and M{\"u}ller, Marina E. H.}, title = {Spotlight on the Underdogs-An Analysis of Underrepresented Alternaria Mycotoxins Formed Depending on Varying Substrate, Time and Temperature Conditions}, series = {Toxins}, volume = {8}, journal = {Toxins}, publisher = {MDPI}, address = {Basel}, issn = {2072-6651}, doi = {10.3390/toxins8110344}, pages = {570 -- 583}, year = {2016}, abstract = {Alternaria (A.) is a genus of widespread fungi capable of producing numerous, possibly health-endangering Alternaria toxins (ATs), which are usually not the focus of attention. The formation of ATs depends on the species and complex interactions of various environmental factors and is not fully understood. In this study the influence of temperature (7 degrees C, 25 degrees C), substrate (rice, wheat kernels) and incubation time (4, 7, and 14 days) on the production of thirteen ATs and three sulfoconjugated ATs by three different Alternaria isolates from the species groups A. tenuissima and A. infectoria was determined. High-performance liquid chromatography coupled with tandem mass spectrometry was used for quantification. Under nearly all conditions, tenuazonic acid was the most extensively produced toxin. At 25 degrees C and with increasing incubation time all toxins were formed in high amounts by the two A. tenuissima strains on both substrates with comparable mycotoxin profiles. However, for some of the toxins, stagnation or a decrease in production was observed from day 7 to 14. As opposed to the A. tenuissima strains, the A. infectoria strain only produced low amounts of ATs, but high concentrations of stemphyltoxin III. The results provide an essential insight into the quantitative in vitro AT formation under different environmental conditions, potentially transferable to different field and storage conditions.}, language = {en} } @misc{ZwickelKahlKlaffkeetal.2017, author = {Zwickel, Theresa and Kahl, Sandra M. and Klaffke, Horst and Rychlik, Michael and M{\"u}ller, Marina E. H.}, title = {Spotlight on the underdogs}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus4-400438}, pages = {17}, year = {2017}, abstract = {Alternaria (A.) is a genus of widespread fungi capable of producing numerous, possibly health-endangering Alternaria toxins (ATs), which are usually not the focus of attention. The formation of ATs depends on the species and complex interactions of various environmental factors and is not fully understood. In this study the influence of temperature (7 °C, 25 °C), substrate (rice, wheat kernels) and incubation time (4, 7, and 14 days) on the production of thirteen ATs and three sulfoconjugated ATs by three different Alternaria isolates from the species groups A. tenuissima and A. infectoria was determined. High-performance liquid chromatography coupled with tandem mass spectrometry was used for quantification. Under nearly all conditions, tenuazonic acid was the most extensively produced toxin. At 25 °C and with increasing incubation time all toxins were formed in high amounts by the two A. tenuissima strains on both substrates with comparable mycotoxin profiles. However, for some of the toxins, stagnation or a decrease in production was observed from day 7 to 14. As opposed to the A. tenuissima strains, the A. infectoria strain only produced low amounts of ATs, but high concentrations of stemphyltoxin III. The results provide an essential insight into the quantitative in vitro AT formation under different environmental conditions, potentially transferable to different field and storage conditions}, language = {en} } @misc{ZwickelKahlRychliketal.2018, author = {Zwickel, Theresa and Kahl, Sandra M. and Rychlik, Michael and M{\"u}ller, Marina E. H.}, title = {Chemotaxonomy of mycotoxigenic small-spored Alternaria fungi}, series = {Postprints der Universit{\"a}t Potsdam Mathematisch-Naturwissenschaftliche Reihe}, journal = {Postprints der Universit{\"a}t Potsdam Mathematisch-Naturwissenschaftliche Reihe}, number = {696}, issn = {1866-8372}, doi = {10.25932/publishup-42662}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus4-426623}, pages = {20}, year = {2018}, abstract = {Necrotrophic as well as saprophytic small-spored Altemaria (A.) species are annually responsible for major losses of agricultural products, such as cereal crops, associated with the contamination of food and feedstuff with potential health-endangering Altemaria toxins. Knowledge of the metabolic capabilities of different species-groups to form mycotoxins is of importance for a reliable risk assessment. 93 Altemaria strains belonging to the four species groups Alternaria tenuissima, A. arborescens, A. altemata, and A. infectoria were isolated from winter wheat kernels harvested from fields in Germany and Russia and incubated under equal conditions. Chemical analysis by means of an HPLC-MS/MS multi-Alternaria-toxin-method showed that 95\% of all strains were able to form at least one of the targeted 17 non-host specific Altemaria toxins. Simultaneous production of up to 15 (modified) Altemaria toxins by members of the A. tenuissima, A. arborescens, A. altemata species-groups and up to seven toxins by A. infectoria strains was demonstrated. Overall tenuazonic acid was the most extensively formed mycotoxin followed by alternariol and alternariol mono methylether, whereas altertoxin I was the most frequently detected toxin. Sulfoconjugated modifications of alternariol, alternariol mono methylether, altenuisol and altenuene were frequently determined. Unknown perylene quinone derivatives were additionally detected. Strains of the species-group A. infectoria could be segregated from strains of the other three species-groups due to significantly lower toxin levels and the specific production of infectopyrone. Apart from infectopyrone, alterperylenol was also frequently produced by 95\% of the A. infectoria strains. Neither by the concentration nor by the composition of the targeted Altemaria toxins a differentiation between the species-groups A. altemata, A. tenuissima and A. arborescens was possible.}, language = {en} } @misc{RubeyBruneHeineetal.2017, author = {Rubey, Michael and Brune, Sascha and Heine, Christian and Davies, D. Rhodri and Williams, Simon E. and M{\"u}ller, R. Dietmar}, title = {Global patterns in Earth's dynamic topography since the Jurassic}, series = {Postprints der Universit{\"a}t Potsdam : Mathematisch Naturwissenschaftliche Reihe}, journal = {Postprints der Universit{\"a}t Potsdam : Mathematisch Naturwissenschaftliche Reihe}, number = {623}, issn = {1866-8372}, doi = {10.25932/publishup-41824}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus4-418241}, pages = {899 -- 919}, year = {2017}, abstract = {We evaluate the spatial and temporal evolution of Earth's long-wavelength surface dynamic topography since the Jurassic using a series of high-resolution global mantle convection models. These models are Earth-like in terms of convective vigour, thermal structure, surface heat-flux and the geographic distribution of heterogeneity. The models generate a degree-2-dominated spectrum of dynamic topography with negative amplitudes above subducted slabs (i.e. circum-Pacific regions and southern Eurasia) and positive amplitudes elsewhere (i.e. Africa, north-western Eurasia and the central Pacific). Model predictions are compared with published observations and subsidence patterns from well data, both globally and for the Australian and southern African regions. We find that our models reproduce the long-wavelength component of these observations, although observed smaller-scale variations are not reproduced. We subsequently define "geodynamic rules" for how different surface tectonic settings are affected by mantle processes: (i) locations in the vicinity of a subduction zone show large negative dynamic topography amplitudes; (ii) regions far away from convergent margins feature long-term positive dynamic topography; and (iii) rapid variations in dynamic support occur along the margins of overriding plates (e.g. the western US) and at points located on a plate that rapidly approaches a subduction zone (e.g. India and the Arabia Peninsula). Our models provide a predictive quantitative framework linking mantle convection with plate tectonics and sedimentary basin evolution, thus improving our understanding of how subduction and mantle convection affect the spatio-temporal evolution of basin architecture.}, language = {en} } @article{BanksNishiyamaHasebeetal.2011, author = {Banks, Jo Ann and Nishiyama, Tomoaki and Hasebe, Mitsuyasu and Bowman, John L. and Gribskov, Michael and dePamphilis, Claude and Albert, Victor A. and Aono, Naoki and Aoyama, Tsuyoshi and Ambrose, Barbara A. and Ashton, Neil W. and Axtell, Michael J. and Barker, Elizabeth and Barker, Michael S. and Bennetzen, Jeffrey L. and Bonawitz, Nicholas D. and Chapple, Clint and Cheng, Chaoyang and Correa, Luiz Gustavo Guedes and Dacre, Michael and DeBarry, Jeremy and Dreyer, Ingo and Elias, Marek and Engstrom, Eric M. and Estelle, Mark and Feng, Liang and Finet, Cedric and Floyd, Sandra K. and Frommer, Wolf B. and Fujita, Tomomichi and Gramzow, Lydia and Gutensohn, Michael and Harholt, Jesper and Hattori, Mitsuru and Heyl, Alexander and Hirai, Tadayoshi and Hiwatashi, Yuji and Ishikawa, Masaki and Iwata, Mineko and Karol, Kenneth G. and Koehler, Barbara and Kolukisaoglu, Uener and Kubo, Minoru and Kurata, Tetsuya and Lalonde, Sylvie and Li, Kejie and Li, Ying and Litt, Amy and Lyons, Eric and Manning, Gerard and Maruyama, Takeshi and Michael, Todd P. and Mikami, Koji and Miyazaki, Saori and Morinaga, Shin-ichi and Murata, Takashi and M{\"u}ller-R{\"o}ber, Bernd and Nelson, David R. and Obara, Mari and Oguri, Yasuko and Olmstead, Richard G. and Onodera, Naoko and Petersen, Bent Larsen and Pils, Birgit and Prigge, Michael and Rensing, Stefan A. and Mauricio Riano-Pachon, Diego and Roberts, Alison W. and Sato, Yoshikatsu and Scheller, Henrik Vibe and Schulz, Burkhard and Schulz, Christian and Shakirov, Eugene V. and Shibagaki, Nakako and Shinohara, Naoki and Shippen, Dorothy E. and Sorensen, Iben and Sotooka, Ryo and Sugimoto, Nagisa and Sugita, Mamoru and Sumikawa, Naomi and Tanurdzic, Milos and Theissen, Guenter and Ulvskov, Peter and Wakazuki, Sachiko and Weng, Jing-Ke and Willats, William W. G. T. and Wipf, Daniel and Wolf, Paul G. and Yang, Lixing and Zimmer, Andreas D. and Zhu, Qihui and Mitros, Therese and Hellsten, Uffe and Loque, Dominique and Otillar, Robert and Salamov, Asaf and Schmutz, Jeremy and Shapiro, Harris and Lindquist, Erika and Lucas, Susan and Rokhsar, Daniel and Grigoriev, Igor V.}, title = {The selaginella genome identifies genetic changes associated with the evolution of vascular plants}, series = {Science}, volume = {332}, journal = {Science}, number = {6032}, publisher = {American Assoc. for the Advancement of Science}, address = {Washington}, issn = {0036-8075}, doi = {10.1126/science.1203810}, pages = {960 -- 963}, year = {2011}, abstract = {Vascular plants appeared similar to 410 million years ago, then diverged into several lineages of which only two survive: the euphyllophytes (ferns and seed plants) and the lycophytes. We report here the genome sequence of the lycophyte Selaginella moellendorffii (Selaginella), the first nonseed vascular plant genome reported. By comparing gene content in evolutionarily diverse taxa, we found that the transition from a gametophyte- to a sporophyte-dominated life cycle required far fewer new genes than the transition from a nonseed vascular to a flowering plant, whereas secondary metabolic genes expanded extensively and in parallel in the lycophyte and angiosperm lineages. Selaginella differs in posttranscriptional gene regulation, including small RNA regulation of repetitive elements, an absence of the trans-acting small interfering RNA pathway, and extensive RNA editing of organellar genes.}, language = {en} } @article{GomezMerinoBrearleyOrnatowskaetal.2004, author = {Gomez-Merino, Fernando Carlos and Brearley, C. A. and Ornatowska, Magdalena and Abdel-Haliem, Mahmoud E. F. and Zanor, Maria Ines and M{\"u}ller-R{\"o}ber, Bernd}, title = {AtDGK2, a novel diacylglycerol kinase from Arabidopsis thaliana, phosphorylates 1-stearoyl-2-arachidonoyl-sn- glycerol and 1,2-dioleoyl-sn-glycerol and exhibits cold-inducible gene expression}, issn = {0021-9258}, year = {2004}, abstract = {Diacylglycerol kinase (DGK) phosphorylates diacylglycerol (DAG) to generate phosphatidic acid (PA). Both DAG and PA are implicated in signal transduction pathways. DGKs have been widely studied in animals, but their analysis in plants is fragmentary. Here, we report the cloning and biochemical characterization of AtDGK2, encoding DGK from Arabidopsis thaliana. AtDGK2 has a predicted molecular mass of 79.4 kDa and, like AtDGK1 previously reported, harbors two copies of a phorbol ester/DAG-binding domain in its N-terminal region. AtDGK2 belongs to a family of seven DGK genes in A. thaliana. AtDGK3 to AtDGK7 encode similar to55-kDa DGKs that lack a typical phorbol ester/DAG-binding domain. Phylogenetically, plant DGKs fall into three clusters. Members of all three clusters are widely expressed in vascular plants. Recombinant AtDGK2 was expressed in Escherichia coli and biochemically characterized. The enzyme phosphorylated 1,2-dioleoyl-sn-glycerol to yield PA, exhibiting Michaelis-Menten type kinetics. Estimated K-m and V-max values were 125 muM for DAG and 0.25 pmol of PA min(-1) mug(-1), respectively. The enzyme was maximally active at pH 7.2. Its activity was Mg2+-dependent and affected by the presence of detergents, salts, and the DGK inhibitor R59022, but not by Ca2+. AtDGK2 exhibited substrate preference for unsaturated DAG analogues (i.e. 1-stearoyl-2-arachidonoyl-sn-glycerol and 1,2- dioleoyl-sn-glycerol). The AtDGK2 gene is expressed in various tissues of the Arabidopsis plant, including leaves, roots, and flowers, as shown by Northern blot analysis and promoter-reporter gene fusions. We found that AtDGK2 is induced by exposure to low temperature (4degreesC), pointing to a role in cold signal transduction}, language = {en} } @article{ScarpeciZanorMuellerRoeberetal.2013, author = {Scarpeci, Telma E. and Zanor, Maria I. and M{\"u}ller-R{\"o}ber, Bernd and Valle, Estela M.}, title = {Overexpression of AtWRKY30 enhances abiotic stress tolerance during early growth stages in Arabidopsis thaliana}, series = {PLANT MOLECULAR BIOLOGY}, volume = {83}, journal = {PLANT MOLECULAR BIOLOGY}, number = {3}, publisher = {SPRINGER}, address = {DORDRECHT}, issn = {0167-4412}, doi = {10.1007/s11103-013-0090-8}, pages = {265 -- 277}, year = {2013}, abstract = {AtWRKY30 belongs to a higher plant transcription factor superfamily, which responds to pathogen attack. In previous studies, the AtWRKY30 gene was found to be highly and rapidly induced in Arabidopsis thaliana leaves after oxidative stress treatment. In this study, electrophoretic mobility shift assays showed that AtWRKY30 binds with high specificity and affinity to the WRKY consensus sequence (W-box), and also to its own promoter. Analysis of the AtWRKY30 expression pattern by qPCR and using transgenic Arabidopsis lines carrying AtWRKY30 promoter-beta-glucuronidase fusions showed transcriptional activity in leaves subjected to biotic or abiotic stress. Transgenic Arabidopsis plants constitutively overexpressing AtWRKY30 (35S::W30 lines) were more tolerant than wild-type plants to oxidative and salinity stresses during seed germination. The results presented here show that AtWRKY30 is responsive to several stress conditions either from abiotic or biotic origin, suggesting that AtWRKY30 could have a role in the activation of defence responses at early stages of Arabidopsis growth by binding to W-boxes found in promoters of many stress/developmentally regulated genes.}, language = {en} } @article{JungingerKuebelSchacheretal.2013, author = {Junginger, Mathias and K{\"u}bel, Christian and Schacher, Felix H. and M{\"u}ller, Axel H. E. and Taubert, Andreas}, title = {Crystal structure and chemical composition of biomimetric calcium phosphate nanofibers}, doi = {10.1039/c3ra23348k}, year = {2013}, abstract = {Calcium phosphate nanofibers with a diameter of only a few nanometers and a cotton-ball-like aggregate morphology have been reported several times in the literature. Although fiber formation seems reproducible in a variety of conditions, the crystal structure and chemical composition of the fibers have been elusive. Using scanning transmission electron microscopy, low dose electron (nano)diffraction, energy-dispersive X-ray spectroscopy, and energy- filtered transmission electron microscopy, we have assigned crystal structures and chemical compositions to the fibers. Moreover, we demonstrate that the mineralization process yields true polymer/calcium phosphate hybrid materials where the block copolymer template is closely associated with the calcium phosphate.}, language = {en} } @article{JungingerKuebelSchacheretal.2013, author = {Junginger, Mathias and K{\"u}bel, Christian and Schacher, Felix H. and M{\"u}ller, Axel H. E. and Taubert, Andreas}, title = {Crystal structure and chemical composition of biomimetic calcium phosphate nanofibers}, series = {RSC Advances}, volume = {3}, journal = {RSC Advances}, number = {28}, publisher = {Royal Society of Chemistry}, address = {Cambridge}, issn = {2046-2069}, doi = {10.1039/c3ra23348k}, pages = {11301 -- 11308}, year = {2013}, abstract = {Calcium phosphate nanofibers with a diameter of only a few nanometers and a cotton-ball-like aggregate morphology have been reported several times in the literature. Although fiber formation seems reproducible in a variety of conditions, the crystal structure and chemical composition of the fibers have been elusive. Using scanning transmission electron microscopy, low dose electron (nano) diffraction, energy-dispersive X-ray spectroscopy, and energy-filtered transmission electron microscopy, we have assigned crystal structures and chemical compositions to the fibers. Moreover, we demonstrate that the mineralization process yields true polymer/calcium phosphate hybrid materials where the block copolymer template is closely associated with the calcium phosphate.}, language = {en} } @article{RaatzBacchiPirhoferWalzletal.2019, author = {Raatz, Larissa and Bacchi, Nina and Pirhofer Walzl, Karin and Glemnitz, Michael and M{\"u}ller, Marina E. H. and Jasmin Radha, Jasmin and Scherber, Christoph}, title = {How much do we really lose?}, series = {Ecology and Evolution}, volume = {9}, journal = {Ecology and Evolution}, number = {13}, publisher = {John Wiley \& Sons}, address = {S.I.}, issn = {2045-7758}, doi = {10.1002/ece3.5370}, pages = {7838 -- 7848}, year = {2019}, abstract = {Natural landscape elements (NLEs) in agricultural landscapes contribute to biodiversity and ecosystem services, but are also regarded as an obstacle for large-scale agricultural production. However, the effects of NLEs on crop yield have rarely been measured. Here, we investigated how different bordering structures, such as agricultural roads, field-to-field borders, forests, hedgerows, and kettle holes, influence agricultural yields. We hypothesized that (a) yield values at field borders differ from mid-field yields and that (b) the extent of this change in yields depends on the bordering structure. We measured winter wheat yields along transects with log-scaled distances from the border into the agricultural field within two intensively managed agricultural landscapes in Germany (2014 near G{\"o}ttingen, and 2015-2017 in the Uckermark). We observed a yield loss adjacent to every investigated bordering structure of 11\%-38\% in comparison with mid-field yields. However, depending on the bordering structure, this yield loss disappeared at different distances. While the proximity of kettle holes did not affect yields more than neighboring agricultural fields, woody landscape elements had strong effects on winter wheat yields. Notably, 95\% of mid-field yields could already be reached at a distance of 11.3 m from a kettle hole and at a distance of 17.8 m from hedgerows as well as forest borders. Our findings suggest that yield losses are especially relevant directly adjacent to woody landscape elements, but not adjacent to in-field water bodies. This highlights the potential to simultaneously counteract yield losses close to the field border and enhance biodiversity by combining different NLEs in agricultural landscapes such as creating strips of extensive grassland vegetation between woody landscape elements and agricultural fields. In conclusion, our results can be used to quantify ecocompensations to find optimal solutions for the delivery of productive and regulative ecosystem services in heterogeneous agricultural landscapes.}, language = {en} } @article{RaatzBacchiWalzletal.2019, author = {Raatz, Larissa and Bacchi, Nina and Walzl, Karin Pirhofer and Glemnitz, Michael and M{\"u}ller, Marina E. H. and Jasmin Radha, Jasmin and Scherber, Christoph}, title = {How much do we really lose?}, series = {Ecology and evolution}, volume = {9}, journal = {Ecology and evolution}, number = {13}, publisher = {Wiley}, address = {Hoboken}, issn = {2045-7758}, doi = {10.1002/ece3.5370}, pages = {7838 -- 7848}, year = {2019}, abstract = {Natural landscape elements (NLEs) in agricultural landscapes contribute to biodiversity and ecosystem services, but are also regarded as an obstacle for large-scale agricultural production. However, the effects of NLEs on crop yield have rarely been measured. Here, we investigated how different bordering structures, such as agricultural roads, field-to-field borders, forests, hedgerows, and kettle holes, influence agricultural yields. We hypothesized that (a) yield values at field borders differ from mid-field yields and that (b) the extent of this change in yields depends on the bordering structure. We measured winter wheat yields along transects with log-scaled distances from the border into the agricultural field within two intensively managed agricultural landscapes in Germany (2014 near Gottingen, and 2015-2017 in the Uckermark). We observed a yield loss adjacent to every investigated bordering structure of 11\%-38\% in comparison with mid-field yields. However, depending on the bordering structure, this yield loss disappeared at different distances. While the proximity of kettle holes did not affect yields more than neighboring agricultural fields, woody landscape elements had strong effects on winter wheat yields. Notably, 95\% of mid-field yields could already be reached at a distance of 11.3 m from a kettle hole and at a distance of 17.8 m from hedgerows as well as forest borders. Our findings suggest that yield losses are especially relevant directly adjacent to woody landscape elements, but not adjacent to in-field water bodies. This highlights the potential to simultaneously counteract yield losses close to the field border and enhance biodiversity by combining different NLEs in agricultural landscapes such as creating strips of extensive grassland vegetation between woody landscape elements and agricultural fields. In conclusion, our results can be used to quantify ecocompensations to find optimal solutions for the delivery of productive and regulative ecosystem services in heterogeneous agricultural landscapes.}, language = {en} } @misc{JungingerKuebelSchacheretal.2013, author = {Junginger, Mathias and K{\"u}bel, Christian and Schacher, Felix H. and M{\"u}ller, Axel H. E. and Taubert, Andreas}, title = {Crystal structure and chemical composition of biomimetic calcium phosphate nanofibers}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus4-95176}, pages = {11301 -- 11308}, year = {2013}, abstract = {Calcium phosphate nanofibers with a diameter of only a few nanometers and a cotton-ball-like aggregate morphology have been reported several times in the literature. Although fiber formation seems reproducible in a variety of conditions, the crystal structure and chemical composition of the fibers have been elusive. Using scanning transmission electron microscopy, low dose electron (nano)diffraction, energy-dispersive X-ray spectroscopy, and energy-filtered transmission electron microscopy, we have assigned crystal structures and chemical compositions to the fibers. Moreover, we demonstrate that the mineralization process yields true polymer/calcium phosphate hybrid materials where the block copolymer template is closely associated with the calcium phosphate.}, language = {en} } @article{MartielMuellerWerkmeisterCohen2019, author = {Martiel, Isabelle and M{\"u}ller-Werkmeister, Henrike and Cohen, Aina E.}, title = {Strategies for sample delivery for femtosecond crystallography}, series = {Acta Crystallographica : Section D, Structural biology}, volume = {75}, journal = {Acta Crystallographica : Section D, Structural biology}, publisher = {Bognor Regis}, address = {Wiley}, issn = {2059-7983}, doi = {10.1107/S2059798318017953}, pages = {160 -- 177}, year = {2019}, abstract = {Highly efficient data-collection methods are required for successful macromolecular crystallography (MX) experiments at X-ray free-electron lasers (XFELs). XFEL beamtime is scarce, and the high peak brightness of each XFEL pulse destroys the exposed crystal volume. It is therefore necessary to combine diffraction images from a large number of crystals (hundreds to hundreds of thousands) to obtain a final data set, bringing about sample-refreshment challenges that have previously been unknown to the MX synchrotron community. In view of this experimental complexity, a number of sample delivery methods have emerged, each with specific requirements, drawbacks and advantages. To provide useful selection criteria for future experiments, this review summarizes the currently available sample delivery methods, emphasising the basic principles and the specific sample requirements. Two main approaches to sample delivery are first covered: (i) injector methods with liquid or viscous media and (ii) fixed-target methods using large crystals or using microcrystals inside multi-crystal holders or chips. Additionally, hybrid methods such as acoustic droplet ejection and crystal extraction are covered, which combine the advantages of both fixed-target and injector approaches.}, language = {en} } @misc{RaatzBacchiPirhoferWalzletal.2019, author = {Raatz, Larissa and Bacchi, Nina and Pirhofer Walzl, Karin and Glemnitz, Michael and M{\"u}ller, Marina E. H. and Jasmin Radha, Jasmin and Scherber, Christoph}, title = {How much do we really lose?}, series = {Postprints der Universit{\"a}t Potsdam : Mathematisch-Naturwissenschaftliche Reihe}, journal = {Postprints der Universit{\"a}t Potsdam : Mathematisch-Naturwissenschaftliche Reihe}, number = {811}, issn = {1866-8372}, doi = {10.25932/publishup-44331}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus4-443313}, pages = {13}, year = {2019}, abstract = {Natural landscape elements (NLEs) in agricultural landscapes contribute to biodiversity and ecosystem services, but are also regarded as an obstacle for large-scale agricultural production. However, the effects of NLEs on crop yield have rarely been measured. Here, we investigated how different bordering structures, such as agricultural roads, field-to-field borders, forests, hedgerows, and kettle holes, influence agricultural yields. We hypothesized that (a) yield values at field borders differ from mid-field yields and that (b) the extent of this change in yields depends on the bordering structure. We measured winter wheat yields along transects with log-scaled distances from the border into the agricultural field within two intensively managed agricultural landscapes in Germany (2014 near G{\"o}ttingen, and 2015-2017 in the Uckermark). We observed a yield loss adjacent to every investigated bordering structure of 11\%-38\% in comparison with mid-field yields. However, depending on the bordering structure, this yield loss disappeared at different distances. While the proximity of kettle holes did not affect yields more than neighboring agricultural fields, woody landscape elements had strong effects on winter wheat yields. Notably, 95\% of mid-field yields could already be reached at a distance of 11.3 m from a kettle hole and at a distance of 17.8 m from hedgerows as well as forest borders. Our findings suggest that yield losses are especially relevant directly adjacent to woody landscape elements, but not adjacent to in-field water bodies. This highlights the potential to simultaneously counteract yield losses close to the field border and enhance biodiversity by combining different NLEs in agricultural landscapes such as creating strips of extensive grassland vegetation between woody landscape elements and agricultural fields. In conclusion, our results can be used to quantify ecocompensations to find optimal solutions for the delivery of productive and regulative ecosystem services in heterogeneous agricultural landscapes.}, language = {en} }