@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} } @misc{Mueller1983, author = {M{\"u}ller, Richard G. E.}, title = {Groffmann, K. J. ; Michel, L. (Hrsg.): Grundlagen psychologischer Diagnostik. - (Enzyklop{\"a}die der Psychologie ; 1) / [rezensiert von] Richard G. E. M{\"u}ller}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-59504}, year = {1983}, abstract = {rezensiertes Werk: Groffmann, K. J. ; Michel, L. (Hrsg.): Grundlagen psychologischer Diagnostik. - G{\"o}ttingen [u.a.] : Hogrefe, Verl. f{\"u}r Psychologie, 1982. - 602 S.. - (Enzyklop{\"a}die der Psychologie ; 1) ISBN: 3-8017-0502-1}, language = {de} } @article{AllanBossdorfDormannetal.2014, author = {Allan, Eric and Bossdorf, Oliver and Dormann, Carsten F. and Prati, Daniel and Gossner, Martin M. and Tscharntke, Teja and Bl{\"u}thgen, Nico and Bellach, Michaela and Birkhofer, Klaus and Boch, Steffen and B{\"o}hm, Stefan and B{\"o}rschig, Carmen and Chatzinotas, Antonis and Christ, Sabina and Daniel, Rolf and Diek{\"o}tter, Tim and Fischer, Christiane and Friedl, Thomas and Glaser, Karin and Hallmann, Christine and Hodac, Ladislav and H{\"o}lzel, Norbert and Jung, Kirsten and Klein, Alexandra-Maria and Klaus, Valentin H. and Kleinebecker, Till and Krauss, Jochen and Lange, Markus and Morris, E. Kathryn and M{\"u}ller, J{\"o}rg and Nacke, Heiko and Pasalic, Esther and Rillig, Matthias C. and Rothenwoehrer, Christoph and Schally, Peter and Scherber, Christoph and Schulze, Waltraud X. and Socher, Stephanie A. and Steckel, Juliane and Steffan-Dewenter, Ingolf and T{\"u}rke, Manfred and Weiner, Christiane N. and Werner, Michael and Westphal, Catrin and Wolters, Volkmar and Wubet, Tesfaye and Gockel, Sonja and Gorke, Martin and Hemp, Andreas and Renner, Swen C. and Sch{\"o}ning, Ingo and Pfeiffer, Simone and K{\"o}nig-Ries, Birgitta and Buscot, Francois and Linsenmair, Karl Eduard and Schulze, Ernst-Detlef and Weisser, Wolfgang W. and Fischer, Markus}, title = {Interannual variation in land-use intensity enhances grassland multidiversity}, series = {Proceedings of the National Academy of Sciences of the United States of America}, volume = {111}, journal = {Proceedings of the National Academy of Sciences of the United States of America}, number = {1}, publisher = {National Acad. of Sciences}, address = {Washington}, issn = {0027-8424}, doi = {10.1073/pnas.1312213111}, pages = {308 -- 313}, year = {2014}, abstract = {Although temporal heterogeneity is a well-accepted driver of biodiversity, effects of interannual variation in land-use intensity (LUI) have not been addressed yet. Additionally, responses to land use can differ greatly among different organisms; therefore, overall effects of land-use on total local biodiversity are hardly known. To test for effects of LUI (quantified as the combined intensity of fertilization, grazing, and mowing) and interannual variation in LUI (SD in LUI across time), we introduce a unique measure of whole-ecosystem biodiversity, multidiversity. This synthesizes individual diversity measures across up to 49 taxonomic groups of plants, animals, fungi, and bacteria from 150 grasslands. Multidiversity declined with increasing LUI among grasslands, particularly for rarer species and aboveground organisms, whereas common species and belowground groups were less sensitive. However, a high level of interannual variation in LUI increased overall multidiversity at low LUI and was even more beneficial for rarer species because it slowed the rate at which the multidiversity of rare species declined with increasing LUI. In more intensively managed grasslands, the diversity of rarer species was, on average, 18\% of the maximum diversity across all grasslands when LUI was static over time but increased to 31\% of the maximum when LUI changed maximally over time. In addition to decreasing overall LUI, we suggest varying LUI across years as a complementary strategy to promote biodiversity conservation.}, language = {en} } @article{LotkowskaTohgeFernieetal.2015, author = {Lotkowska, Magda E. and Tohge, Takayuki and Fernie, Alisdair R. and Xue, Gang-Ping and Balazadeh, Salma and M{\"u}ller-R{\"o}ber, Bernd}, title = {The Arabidopsis Transcription Factor MYB112 Promotes Anthocyanin Formation during Salinity and under High Light Stress}, series = {Plant physiology : an international journal devoted to physiology, biochemistry, cellular and molecular biology, biophysics and environmental biology of plants}, volume = {169}, journal = {Plant physiology : an international journal devoted to physiology, biochemistry, cellular and molecular biology, biophysics and environmental biology of plants}, number = {3}, publisher = {American Society of Plant Physiologists}, address = {Rockville}, issn = {0032-0889}, doi = {10.1104/pp.15.00605}, pages = {1862 -- 1880}, year = {2015}, abstract = {MYB transcription factors (TFs) are important regulators of flavonoid biosynthesis in plants. Here, we report MYB112 as a formerly unknown regulator of anthocyanin accumulation in Arabidopsis (Arabidopsis thaliana). Expression profiling after chemically induced overexpression of MYB112 identified 28 up-and 28 down-regulated genes 5 h after inducer treatment, including MYB7 and MYB32, which are both induced. In addition, upon extended induction, MYB112 also positively affects the expression of PRODUCTION OF ANTHOCYANIN PIGMENT1, a key TF of anthocyanin biosynthesis, but acts negatively toward MYB12 and MYB111, which both control flavonol biosynthesis. MYB112 binds to an 8-bp DNA fragment containing the core sequence (A/T/G)(A/C) CC(A/T)(A/G/T)(A/C)(T/C). By electrophoretic mobility shift assay and chromatin immunoprecipitation coupled to quantitative polymerase chain reaction, we show that MYB112 binds in vitro and in vivo to MYB7 and MYB32 promoters, revealing them as direct downstream target genes. We further show that MYB112 expression is up-regulated by salinity and high light stress, environmental parameters that both require the MYB112 TF for anthocyanin accumulation under these stresses. In contrast to several other MYB TFs affecting anthocyanin biosynthesis, MYB112 expression is not controlled by nitrogen limitation or an excess of carbon. Thus, MYB112 constitutes a regulator that promotes anthocyanin accumulation under abiotic stress conditions.}, language = {en} } @article{RaatzPirhoferWalzlMuelleretal.2021, author = {Raatz, Larissa and Pirhofer-Walzl, Karin and M{\"u}ller, Marina E.H. and Scherber, Christoph and Joshi, Jasmin Radha}, title = {Who is the culprit: Is pest infestation responsible for crop yield losses close to semi-natural habitats?}, series = {Ecology and Evolution}, volume = {11}, journal = {Ecology and Evolution}, edition = {19}, publisher = {Wiley-Blackwell}, address = {Oxford}, issn = {1467-6435}, doi = {10.1002/ece3.8046}, pages = {13232 -- 13246}, year = {2021}, abstract = {Semi-natural habitats (SNHs) are becoming increasingly scarce in modern agricultural landscapes. This may reduce natural ecosystem services such as pest control with its putatively positive effect on crop production. In agreement with other studies, we recently reported wheat yield reductions at field borders which were linked to the type of SNH and the distance to the border. In this experimental landscape-wide study, we asked whether these yield losses have a biotic origin while analyzing fungal seed and fungal leaf pathogens, herbivory of cereal leaf beetles, and weed cover as hypothesized mediators between SNHs and yield. We established experimental winter wheat plots of a single variety within conventionally managed wheat fields at fixed distances either to a hedgerow or to an in-field kettle hole. For each plot, we recorded the fungal infection rate on seeds, fungal infection and herbivory rates on leaves, and weed cover. Using several generalized linear mixed-effects models as well as a structural equation model, we tested the effects of SNHs at a field scale (SNH type and distance to SNH) and at a landscape scale (percentage and diversity of SNHs within a 1000-m radius). In the dry year of 2016, we detected one putative biotic culprit: Weed cover was negatively associated with yield values at a 1-m and 5-m distance from the field border with a SNH. None of the fungal and insect pests, however, significantly affected yield, neither solely nor depending on type of or distance to a SNH. However, the pest groups themselves responded differently to SNH at the field scale and at the landscape scale. Our findings highlight that crop losses at field borders may be caused by biotic culprits; however, their negative impact seems weak and is putatively reduced by conventional farming practices.}, language = {en} } @article{RaatzPirhoferWalzlMuelleretal.2021, author = {Raatz, Larissa and Pirhofer-Walzl, Karin and M{\"u}ller, Marina E.H. and Scherber, Christoph and Joshi, Jasmin Radha}, title = {Who is the culprit: Is pest infestation responsible for crop yield losses close to semi-natural habitats?}, series = {Zweitver{\"o}ffentlichungen der Universit{\"a}t Potsdam : Mathematisch-Naturwissenschaftliche Reihe}, journal = {Zweitver{\"o}ffentlichungen der Universit{\"a}t Potsdam : Mathematisch-Naturwissenschaftliche Reihe}, publisher = {Universit{\"a}tsverlag Potsdam}, address = {Potsdam}, issn = {1866-8372}, doi = {10.25932/publishup-54962}, url = {http://nbn-resolving.de/urn:nbn:de:kobv:517-opus4-549622}, pages = {13232 -- 13246}, year = {2021}, abstract = {Semi-natural habitats (SNHs) are becoming increasingly scarce in modern agricultural landscapes. This may reduce natural ecosystem services such as pest control with its putatively positive effect on crop production. In agreement with other studies, we recently reported wheat yield reductions at field borders which were linked to the type of SNH and the distance to the border. In this experimental landscape-wide study, we asked whether these yield losses have a biotic origin while analyzing fungal seed and fungal leaf pathogens, herbivory of cereal leaf beetles, and weed cover as hypothesized mediators between SNHs and yield. We established experimental winter wheat plots of a single variety within conventionally managed wheat fields at fixed distances either to a hedgerow or to an in-field kettle hole. For each plot, we recorded the fungal infection rate on seeds, fungal infection and herbivory rates on leaves, and weed cover. Using several generalized linear mixed-effects models as well as a structural equation model, we tested the effects of SNHs at a field scale (SNH type and distance to SNH) and at a landscape scale (percentage and diversity of SNHs within a 1000-m radius). In the dry year of 2016, we detected one putative biotic culprit: Weed cover was negatively associated with yield values at a 1-m and 5-m distance from the field border with a SNH. None of the fungal and insect pests, however, significantly affected yield, neither solely nor depending on type of or distance to a SNH. However, the pest groups themselves responded differently to SNH at the field scale and at the landscape scale. Our findings highlight that crop losses at field borders may be caused by biotic culprits; however, their negative impact seems weak and is putatively reduced by conventional farming practices.}, language = {en} }