@article{GalluzziBravoSanPedroVitaleetal.2015, author = {Galluzzi, L. and Bravo-San Pedro, J. M. and Vitale, I. and Aaronson, S. A. and Abrams, J. M. and Adam, D. and Alnemri, E. S. and Altucci, L. and Andrews, D. and Annicchiarico-Petruzelli, M. and Baehrecke, E. H. and Bazan, N. G. and Bertrand, M. J. and Bianchi, K. and Blagosklonny, M. V. and Blomgren, K. and Borner, C. and Bredesen, D. E. and Brenner, C. and Campanella, M. and Candi, E. and Cecconi, F. and Chan, F. K. and Chandel, N. S. and Cheng, E. H. and Chipuk, J. E. and Cidlowski, J. A. and Ciechanover, A. and Dawson, T. M. and Dawson, V. L. and De Laurenzi, V. and De Maria, R. and Debatin, K. M. and Di Daniele, N. and Dixit, V. M. and Dynlacht, B. D. and El-Deiry, W. S. and Fimia, G. M. and Flavell, R. A. and Fulda, S. and Garrido, C. and Gougeon, M. L. and Green, D. R. and Gronemeyer, H. and Hajnoczky, G. and Hardwick, J. M. and Hengartner, M. O. and Ichijo, H. and Joseph, B. and Jost, P. J. and Kaufmann, T. and Kepp, O. and Klionsky, D. J. and Knight, R. A. and Kumar, S. and Lemasters, J. J. and Levine, B. and Linkermann, A. and Lipton, S. A. and Lockshin, R. A. and L{\´o}pez-Ot{\´i}n, C. and Lugli, E. and Madeo, F. and Malorni, W. and Marine, J. C. and Martin, S. J. and Martinou, J. C. and Medema, J. P. and Meier, P. and Melino, S. and Mizushima, N. and Moll, U. and Mu{\~n}oz-Pinedo, C. and Nu{\~n}ez, G. and Oberst, A. and Panaretakis, T. and Penninger, J. M. and Peter, M. E. and Piacentini, M. and Pinton, P. and Prehn, J. H. and Puthalakath, H. and Rabinovich, G. A. and Ravichandran, K. S. and Rizzuto, R. and Rodrigues, C. M. and Rubinsztein, D. C. and Rudel, T. and Shi, Y. and Simon, H. U. and Stockwell, B. R. and Szabadkai, G. and Tait, S. W. and Tang, H. L. and Tavernarakis, N. and Tsujimoto, Y. and Vanden Berghe, T. and Vandenabeele, P. and Villunger, A. and Wagner, E. F. and Walczak, H. and White, E. and Wood, W. G. and Yuan, J. and Zakeri, Z. and Zhivotovsky, B. and Melino, G. and Kroemer, G.}, title = {Essential versus accessory aspects of cell death: recommendations of the NCCD 2015}, series = {Cell Death and Differentiation}, volume = {22}, journal = {Cell Death and Differentiation}, doi = {10.1038/cdd.2014.137}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-121207}, pages = {58-73}, year = {2015}, abstract = {Cells exposed to extreme physicochemical or mechanical stimuli die in an uncontrollable manner, as a result of their immediate structural breakdown. Such an unavoidable variant of cellular demise is generally referred to as 'accidental cell death' (ACD). In most settings, however, cell death is initiated by a genetically encoded apparatus, correlating with the fact that its course can be altered by pharmacologic or genetic interventions. 'Regulated cell death' (RCD) can occur as part of physiologic programs or can be activated once adaptive responses to perturbations of the extracellular or intracellular microenvironment fail. The biochemical phenomena that accompany RCD may be harnessed to classify it into a few subtypes, which often (but not always) exhibit stereotyped morphologic features. Nonetheless, efficiently inhibiting the processes that are commonly thought to cause RCD, such as the activation of executioner caspases in the course of apoptosis, does not exert true cytoprotective effects in the mammalian system, but simply alters the kinetics of cellular demise as it shifts its morphologic and biochemical correlates. Conversely, bona fide cytoprotection can be achieved by inhibiting the transduction of lethal signals in the early phases of the process, when adaptive responses are still operational. Thus, the mechanisms that truly execute RCD may be less understood, less inhibitable and perhaps more homogeneous than previously thought. Here, the Nomenclature Committee on Cell Death formulates a set of recommendations to help scientists and researchers to discriminate between essential and accessory aspects of cell death.}, language = {en} } @article{ViljurAbellaAdameketal.2022, author = {Viljur, Mari-Liis and Abella, Scott R. and Ad{\´a}mek, Martin and Alencar, Janderson Batista Rodrigues and Barber, Nicholas A. and Beudert, Burkhard and Burkle, Laura A. and Cagnolo, Luciano and Campos, Brent R. and Chao, Anne and Chergui, Brahim and Choi, Chang-Yong and Cleary, Daniel F. R. and Davis, Thomas Seth and Dechnik-V{\´a}zquez, Yanus A. and Downing, William M. and Fuentes-Ramirez, Andr{\´e}s and Gandhi, Kamal J. K. and Gehring, Catherine and Georgiev, Kostadin B. and Gimbutas, Mark and Gongalsky, Konstantin B. and Gorbunova, Anastasiya Y. and Greenberg, Cathryn H. and Hylander, Kristoffer and Jules, Erik S. and Korobushkin, Daniil I. and K{\"o}ster, Kajar and Kurth, Valerie and Lanham, Joseph Drew and Lazarina, Maria and Leverkus, Alexandro B. and Lindenmayer, David and Marra, Daniel Magnabosco and Mart{\´i}n-Pinto, Pablo and Meave, Jorge A. and Moretti, Marco and Nam, Hyun-Young and Obrist, Martin K. and Petanidou, Theodora and Pons, Pere and Potts, Simon G. and Rapoport, Irina B. and Rhoades, Paul R. and Richter, Clark and Saifutdinov, Ruslan A. and Sanders, Nathan J. and Santos, Xavier and Steel, Zachary and Tavella, Julia and Wendenburg, Clara and Wermelinger, Beat and Zaitsev, Andrey S. and Thorn, Simon}, title = {The effect of natural disturbances on forest biodiversity: an ecological synthesis}, series = {Biological Reviews}, volume = {97}, journal = {Biological Reviews}, number = {5}, doi = {10.1111/brv.12876}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-287168}, pages = {1930 -- 1947}, year = {2022}, abstract = {Disturbances alter biodiversity via their specific characteristics, including severity and extent in the landscape, which act at different temporal and spatial scales. Biodiversity response to disturbance also depends on the community characteristics and habitat requirements of species. Untangling the mechanistic interplay of these factors has guided disturbance ecology for decades, generating mixed scientific evidence of biodiversity responses to disturbance. Understanding the impact of natural disturbances on biodiversity is increasingly important due to human-induced changes in natural disturbance regimes. In many areas, major natural forest disturbances, such as wildfires, windstorms, and insect outbreaks, are becoming more frequent, intense, severe, and widespread due to climate change and land-use change. Conversely, the suppression of natural disturbances threatens disturbance-dependent biota. Using a meta-analytic approach, we analysed a global data set (with most sampling concentrated in temperate and boreal secondary forests) of species assemblages of 26 taxonomic groups, including plants, animals, and fungi collected from forests affected by wildfires, windstorms, and insect outbreaks. The overall effect of natural disturbances on α-diversity did not differ significantly from zero, but some taxonomic groups responded positively to disturbance, while others tended to respond negatively. Disturbance was beneficial for taxonomic groups preferring conditions associated with open canopies (e.g. hymenopterans and hoverflies), whereas ground-dwelling groups and/or groups typically associated with shady conditions (e.g. epigeic lichens and mycorrhizal fungi) were more likely to be negatively impacted by disturbance. Across all taxonomic groups, the highest α-diversity in disturbed forest patches occurred under moderate disturbance severity, i.e. with approximately 55\% of trees killed by disturbance. We further extended our meta-analysis by applying a unified diversity concept based on Hill numbers to estimate α-diversity changes in different taxonomic groups across a gradient of disturbance severity measured at the stand scale and incorporating other disturbance features. We found that disturbance severity negatively affected diversity for Hill number q = 0 but not for q = 1 and q = 2, indicating that diversity-disturbance relationships are shaped by species relative abundances. Our synthesis of α-diversity was extended by a synthesis of disturbance-induced change in species assemblages, and revealed that disturbance changes the β-diversity of multiple taxonomic groups, including some groups that were not affected at the α-diversity level (birds and woody plants). Finally, we used mixed rarefaction/extrapolation to estimate biodiversity change as a function of the proportion of forests that were disturbed, i.e. the disturbance extent measured at the landscape scale. The comparison of intact and naturally disturbed forests revealed that both types of forests provide habitat for unique species assemblages, whereas species diversity in the mixture of disturbed and undisturbed forests peaked at intermediate values of disturbance extent in the simulated landscape. Hence, the relationship between α-diversity and disturbance severity in disturbed forest stands was strikingly similar to the relationship between species richness and disturbance extent in a landscape consisting of both disturbed and undisturbed forest habitats. This result suggests that both moderate disturbance severity and moderate disturbance extent support the highest levels of biodiversity in contemporary forest landscapes.}, language = {en} } @article{FerreiraGamazonAlEjehetal.2019, author = {Ferreira, Manuel A. and Gamazon, Eric R. and Al-Ejeh, Fares and Aittom{\"a}ki, Kristiina and Andrulis, Irene L. and Anton-Culver, Hoda and Arason, Adalgeir and Arndt, Volker and Aronson, Kristan J. and Arun, Banu K. and Asseryanis, Ella and Azzollini, Jacopo and Balma{\~n}a, Judith and Barnes, Daniel R. and Barrowdale, Daniel and Beckmann, Matthias W. and Behrens, Sabine and Benitez, Javier and Bermisheva, Marina and Bialkowska, Katarzyna and Blomqvist, Carl and Bogdanova, Natalia V. and Bojesen, Stig E. and Bolla, Manjeet K. and Borg, Ake and Brauch, Hiltrud and Brenner, Hermann and Broeks, Annegien and Burwinkel, Barbara and Cald{\´e}s, Trinidad and Caligo, Maria A. and Campa, Daniele and Campbell, Ian and Canzian, Federico and Carter, Jonathan and Carter, Brian D. and Castelao, Jose E. and Chang-Claude, Jenny and Chanock, Stephen J. and Christiansen, Hans and Chung, Wendy K. and Claes, Kathleen B. M. and Clarke, Christine L. and Couch, Fergus J. and Cox, Angela and Cross, Simon S. and Czene, Kamila and Daly, Mary B. and de la Hoya, Miguel and Dennis, Joe and Devilee, Peter and Diez, Orland and D{\"o}rk, Thilo and Dunning, Alison M. and Dwek, Miriam and Eccles, Diana M. and Ejlertsen, Bent and Ellberg, Carolina and Engel, Christoph and Eriksson, Mikael and Fasching, Peter A. and Fletcher, Olivia and Flyger, Henrik and Friedman, Eitan and Frost, Debra and Gabrielson, Marike and Gago-Dominguez, Manuela and Ganz, Patricia A. and Gapstur, Susan M. and Garber, Judy and Garc{\´i}a-Closas, Montserrat and Garc{\´i}a-S{\´a}enz, Jos{\´e} A. and Gaudet, Mia M. and Giles, Graham G. and Glendon, Gord and Godwin, Andrew K. and Goldberg, Mark S. and Goldgar, David E. and Gonz{\´a}lez-Neira, Anna and Greene, Mark H. and Gronwald, Jacek and Guen{\´e}l, Pascal and Haimann, Christopher A. and Hall, Per and Hamann, Ute and He, Wei and Heyworth, Jane and Hogervorst, Frans B. L. and Hollestelle, Antoinette and Hoover, Robert N. and Hopper, John L. and Hulick, Peter J. and Humphreys, Keith and Imyanitov, Evgeny N. and Isaacs, Claudine and Jakimovska, Milena and Jakubowska, Anna and James, Paul A. and Janavicius, Ramunas and Jankowitz, Rachel C. and John, Esther M. and Johnson, Nichola and Joseph, Vijai and Karlan, Beth Y. and Khusnutdinova, Elza and Kiiski, Johanna I. and Ko, Yon-Dschun and Jones, Michael E. and Konstantopoulou, Irene and Kristensen, Vessela N. and Laitman, Yael and Lambrechts, Diether and Lazaro, Conxi and Leslie, Goska and Lester, Jenny and Lesueur, Fabienne and Lindstr{\"o}m, Sara and Long, Jirong and Loud, Jennifer T. and Lubiński, Jan and Makalic, Enes and Mannermaa, Arto and Manoochehri, Mehdi and Margolin, Sara and Maurer, Tabea and Mavroudis, Dimitrios and McGuffog, Lesley and Meindl, Alfons and Menon, Usha and Michailidou, Kyriaki and Miller, Austin and Montagna, Marco and Moreno, Fernando and Moserle, Lidia and Mulligan, Anna Marie and Nathanson, Katherine L. and Neuhausen, Susan L. and Nevanlinna, Heli and Nevelsteen, Ines and Nielsen, Finn C. and Nikitina-Zake, Liene and Nussbaum, Robert L. and Offit, Kenneth and Olah, Edith and Olopade, Olufunmilayo I. and Olsson, H{\aa}kan and Osorio, Ana and Papp, Janos and Park-Simon, Tjoung-Won and Parsons, Michael T. and Pedersen, Inge Sokilde and Peixoto, Ana and Peterlongo, Paolo and Pharaoh, Paul D. P. and Plaseska-Karanfilska, Dijana and Poppe, Bruce and Presneau, Nadege and Radice, Paolo and Rantala, Johanna and Rennert, Gad and Risch, Harvey A. and Saloustros, Emmanouil and Sanden, Kristin and Sawyer, Elinor J. and Schmidt, Marjanka K. and Schmutzler, Rita K. and Sharma, Priyanka and Shu, Xiao-Ou and Simard, Jaques and Singer, Christian F. and Soucy, Penny and Southey, Melissa C. and Spinelli, John J. and Spurdle, Amanda B. and Stone, Jennifer and Swerdlow, Anthony J. and Tapper, William J. and Taylor, Jack A. and Teixeira, Manuel R. and Terry, Mary Beth and Teul{\´e}, Alex and Thomassen, Mads and Th{\"o}ne, Kathrin and Thull, Darcy L. and Tischkowitz, Marc and Toland, Amanda E. and Torres, Diana and Truong, Th{\´e}r{\`e}se and Tung, Nadine and Vachon, Celine M. and van Asperen, Christi J. and van den Ouweland, Ans M. W. and van Rensburg, Elizabeth J. and Vega, Ana and Viel, Alexandra and Wang, Qin and Wappenschmidt, Barbara and Weitzel, Jeffrey N. and Wendt, Camilla and Winqvist, Robert and Yang, Xiaohong R. and Yannoukakos, Drakoulis and Ziogas, Argyrios and Kraft, Peter and Antoniou, Antonis C. and Zheng, Wei and Easton, Douglas F. and Milne, Roger L. and Beesley, Jonathan and Chenevix-Trench, Georgia}, title = {Genome-wide association and transcriptome studies identify target genes and risk loci for breast cancer}, series = {Nature Communications}, volume = {10}, journal = {Nature Communications}, organization = {EMBRACE Collaborators, GC-HBOC Study Collaborators, GEMO Study Collaborators, ABCTB Investigators, HEBON Investigators, BCFR Investigators}, doi = {10.1038/s41467-018-08053-5}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-228024}, year = {2019}, abstract = {Genome-wide association studies (GWAS) have identified more than 170 breast cancer susceptibility loci. Here we hypothesize that some risk-associated variants might act in non-breast tissues, specifically adipose tissue and immune cells from blood and spleen. Using expression quantitative trait loci (eQTL) reported in these tissues, we identify 26 previously unreported, likely target genes of overall breast cancer risk variants, and 17 for estrogen receptor (ER)-negative breast cancer, several with a known immune function. We determine the directional effect of gene expression on disease risk measured based on single and multiple eQTL. In addition, using a gene-based test of association that considers eQTL from multiple tissues, we identify seven (and four) regions with variants associated with overall (and ER-negative) breast cancer risk, which were not reported in previous GWAS. Further investigation of the function of the implicated genes in breast and immune cells may provide insights into the etiology of breast cancer.}, language = {en} } @article{DavisYuKeenanetal.2013, author = {Davis, Lea K. and Yu, Dongmei and Keenan, Clare L. and Gamazon, Eric R. and Konkashbaev, Anuar I. and Derks, Eske M. and Neale, Benjamin M. and Yang, Jian and Lee, S. Hong and Evans, Patrick and Barr, Cathy L. and Bellodi, Laura and Benarroch, Fortu and Berrio, Gabriel Bedoya and Bienvenu, Oscar J. and Bloch, Michael H. and Blom, Rianne M. and Bruun, Ruth D. and Budman, Cathy L. and Camarena, Beatriz and Campbell, Desmond and Cappi, Carolina and Cardona Silgado, Julio C. and Cath, Danielle C. and Cavallini, Maria C. and Chavira, Denise A. and Chouinard, Sylvian and Conti, David V. and Cook, Edwin H. and Coric, Vladimir and Cullen, Bernadette A. and Deforce, Dieter and Delorme, Richard and Dion, Yves and Edlund, Christopher K. and Egberts, Karin and Falkai, Peter and Fernandez, Thomas V. and Gallagher, Patience J. and Garrido, Helena and Geller, Daniel and Girard, Simon L. and Grabe, Hans J. and Grados, Marco A. and Greenberg, Benjamin D. and Gross-Tsur, Varda and Haddad, Stephen and Heiman, Gary A. and Hemmings, Sian M. J. and Hounie, Ana G. and Illmann, Cornelia and Jankovic, Joseph and Jenike, Micheal A. and Kennedy, James L. and King, Robert A. and Kremeyer, Barbara and Kurlan, Roger and Lanzagorta, Nuria and Leboyer, Marion and Leckman, James F. and Lennertz, Leonhard and Liu, Chunyu and Lochner, Christine and Lowe, Thomas L. and Macciardi, Fabio and McCracken, James T. and McGrath, Lauren M. and Restrepo, Sandra C. Mesa and Moessner, Rainald and Morgan, Jubel and Muller, Heike and Murphy, Dennis L. and Naarden, Allan L. and Ochoa, William Cornejo and Ophoff, Roel A. and Osiecki, Lisa and Pakstis, Andrew J. and Pato, Michele T. and Pato, Carlos N. and Piacentini, John and Pittenger, Christopher and Pollak, Yehunda and Rauch, Scott L. and Renner, Tobias J. and Reus, Victor I. and Richter, Margaret A. and Riddle, Mark A. and Robertson, Mary M. and Romero, Roxana and Ros{\`a}rio, Maria C. and Rosenberg, David and Rouleau, Guy A. and Ruhrmann, Stephan and Ruiz-Linares, Andreas and Sampaio, Aline S. and Samuels, Jack and Sandor, Paul and Sheppard, Broke and Singer, Harvey S. and Smit, Jan H. and Stein, Dan J. and Strengman, E. and Tischfield, Jay A. and Valencia Duarte, Ana V. and Vallada, Homero and Van Nieuwerburgh, Flip and Veenstra-VanderWeele, Jeremy and Walitza, Susanne and Wang, Ying and Wendland, Jens R. and Westenberg, Herman G. M. and Shugart, Yin Yao and Miguel, Euripedes C. and McMahon, William and Wagner, Michael and Nicolini, Humberto and Posthuma, Danielle and Hanna, Gregory L. and Heutink, Peter and Denys, Damiaan and Arnold, Paul D. and Oostra, Ben A. and Nestadt, Gerald and Freimer, Nelson B. and Pauls, David L. and Wray, Naomi R. and Stewart, S. Evelyn and Mathews, Carol A. and Knowles, James A. and Cox, Nancy J. and Scharf, Jeremiah M.}, title = {Partitioning the Heritability of Tourette Syndrome and Obsessive Compulsive Disorder Reveals Differences in Genetic Architecture}, series = {PLoS Genetics}, volume = {9}, journal = {PLoS Genetics}, number = {10}, issn = {1553-7390}, doi = {10.1371/journal.pgen.1003864}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-127377}, pages = {e1003864}, year = {2013}, abstract = {The direct estimation of heritability from genome-wide common variant data as implemented in the program Genome-wide Complex Trait Analysis (GCTA) has provided a means to quantify heritability attributable to all interrogated variants. We have quantified the variance in liability to disease explained by all SNPs for two phenotypically-related neurobehavioral disorders, obsessive-compulsive disorder (OCD) and Tourette Syndrome (TS), using GCTA. Our analysis yielded a heritability point estimate of 0.58 (se = 0.09, p = 5.64e-12) for TS, and 0.37 (se = 0.07, p = 1.5e-07) for OCD. In addition, we conducted multiple genomic partitioning analyses to identify genomic elements that concentrate this heritability. We examined genomic architectures of TS and OCD by chromosome, MAF bin, and functional annotations. In addition, we assessed heritability for early onset and adult onset OCD. Among other notable results, we found that SNPs with a minor allele frequency of less than 5\% accounted for 21\% of the TS heritability and 0\% of the OCD heritability. Additionally, we identified a significant contribution to TS and OCD heritability by variants significantly associated with gene expression in two regions of the brain (parietal cortex and cerebellum) for which we had available expression quantitative trait loci (eQTLs). Finally we analyzed the genetic correlation between TS and OCD, revealing a genetic correlation of 0.41 (se = 0.15, p = 0.002). These results are very close to previous heritability estimates for TS and OCD based on twin and family studies, suggesting that very little, if any, heritability is truly missing (i.e., unassayed) from TS and OCD GWAS studies of common variation. The results also indicate that there is some genetic overlap between these two phenotypically-related neuropsychiatric disorders, but suggest that the two disorders have distinct genetic architectures.}, language = {en} } @article{HudsonNewboldContuetal.2014, author = {Hudson, Lawrence N. and Newbold, Tim and Contu, Sara and Hill, Samantha L. L. and Lysenko, Igor and De Palma, Adriana and Phillips, Helen R. P. and Senior, Rebecca A. and Bennett, Dominic J. and Booth, Hollie and Choimes, Argyrios and Correia, David L. P. and Day, Julie and Echeverria-Londono, Susy and Garon, Morgan and Harrison, Michelle L. K. and Ingram, Daniel J. and Jung, Martin and Kemp, Victoria and Kirkpatrick, Lucinda and Martin, Callum D. and Pan, Yuan and White, Hannah J. and Aben, Job and Abrahamczyk, Stefan and Adum, Gilbert B. and Aguilar-Barquero, Virginia and Aizen, Marcelo and Ancrenaz, Marc and Arbelaez-Cortes, Enrique and Armbrecht, Inge and Azhar, Badrul and Azpiroz, Adrian B. and Baeten, Lander and B{\´a}ldi, Andr{\´a}s and Banks, John E. and Barlow, Jos and Bat{\´a}ry, P{\´e}ter and Bates, Adam J. and Bayne, Erin M. and Beja, Pedro and Berg, Ake and Berry, Nicholas J. and Bicknell, Jake E. and Bihn, Jochen H. and B{\"o}hning-Gaese, Katrin and Boekhout, Teun and Boutin, Celine and Bouyer, Jeremy and Brearley, Francis Q. and Brito, Isabel and Brunet, J{\"o}rg and Buczkowski, Grzegorz and Buscardo, Erika and Cabra-Garcia, Jimmy and Calvino-Cancela, Maria and Cameron, Sydney A. and Cancello, Eliana M. and Carrijo, Tiago F. and Carvalho, Anelena L. and Castro, Helena and Castro-Luna, Alejandro A. and Cerda, Rolando and Cerezo, Alexis and Chauvat, Matthieu and Clarke, Frank M. and Cleary, Daniel F. R. and Connop, Stuart P. and D'Aniello, Biagio and da Silva, Pedro Giovani and Darvill, Ben and Dauber, Jens and Dejean, Alain and Diek{\"o}tter, Tim and Dominguez-Haydar, Yamileth and Dormann, Carsten F. and Dumont, Bertrand and Dures, Simon G. and Dynesius, Mats and Edenius, Lars and Elek, Zolt{\´a}n and Entling, Martin H. and Farwig, Nina and Fayle, Tom M. and Felicioli, Antonio and Felton, Annika M. and Ficetola, Gentile F. and Filgueiras, Bruno K. C. and Fonte, Steve J. and Fraser, Lauchlan H. and Fukuda, Daisuke and Furlani, Dario and Ganzhorn, J{\"o}rg U. and Garden, Jenni G. and Gheler-Costa, Carla and Giordani, Paolo and Giordano, Simonetta and Gottschalk, Marco S. and Goulson, Dave and Gove, Aaron D. and Grogan, James and Hanley, Mick E. and Hanson, Thor and Hashim, Nor R. and Hawes, Joseph E. and H{\´e}bert, Christian and Helden, Alvin J. and Henden, John-Andr{\´e} and Hern{\´a}ndez, Lionel and Herzog, Felix and Higuera-Diaz, Diego and Hilje, Branko and Horgan, Finbarr G. and Horv{\´a}th, Roland and Hylander, Kristoffer and Horv{\´a}th, Roland and Isaacs-Cubides, Paola and Ishitani, Mashiro and Jacobs, Carmen T. and Jaramillo, Victor J. and Jauker, Birgit and Jonsell, Matts and Jung, Thomas S. and Kapoor, Vena and Kati, Vassiliki and Katovai, Eric and Kessler, Michael and Knop, Eva and Kolb, Annette and K{\"o}r{\"o}si, {\`A}d{\´a}m and Lachat, Thibault and Lantschner, Victoria and Le F{\´e}on, Violette and LeBuhn, Gretchen and L{\´e}gar{\´e}, Jean-Philippe and Letcher, Susan G. and Littlewood, Nick A. and L{\´o}pez-Quintero, Carlos A. and Louhaichi, Mounir and L{\"o}vei, Gabor L. and Lucas-Borja, Manuel Esteban and Luja, Victor H. and Maeto, Kaoru and Magura, Tibor and Mallari, Neil Aldrin and Marin-Spiotta, Erika and Marhall, E. J. P. and Mart{\´i}nez, Eliana and Mayfield, Margaret M. and Mikusinski, Gregorz and Milder, Jeffery C. and Miller, James R. and Morales, Carolina L. and Muchane, Mary N. and Muchane, Muchai and Naidoo, Robin and Nakamura, Akihiro and Naoe, Shoji and Nates-Parra, Guiomar and Navarerete Gutierrez, Dario A. and Neuschulz, Eike L. and Noreika, Norbertas and Norfolk, Olivia and Noriega, Jorge Ari and N{\"o}ske, Nicole M. and O'Dea, Niall and Oduro, William and Ofori-Boateng, Caleb and Oke, Chris O. and Osgathorpe, Lynne M. and Paritsis, Juan and Parrah, Alejandro and Pelegrin, Nicol{\´a}s and Peres, Carlos A. and Persson, Anna S. and Petanidou, Theodora and Phalan, Ben and Philips, T. Keith and Poveda, Katja and Power, Eileen F. and Presley, Steven J. and Proen{\c{c}}a, V{\^a}nia and Quaranta, Marino and Quintero, Carolina and Redpath-Downing, Nicola A. and Reid, J. Leighton and Reis, Yana T. and Ribeiro, Danilo B. and Richardson, Barbara A. and Richardson, Michael J. and Robles, Carolina A. and R{\"o}mbke, J{\"o}rg and Romero-Duque, Luz Piedad and Rosselli, Loreta and Rossiter, Stephen J. and Roulston, T'ai H. and Rousseau, Laurent and Sadler, Jonathan P. and S{\´a}fi{\´a}n, Szbolcs and Salda{\~n}a-V{\´a}squez, Romeo A. and Samneg{\aa}rd, Ulrika and Sch{\"u}epp, Christof and Schweiger, Oliver and Sedlock, Jodi L. and Shahabuddin, Ghazala and Sheil, Douglas and Silva, Fernando A. B. and Slade, Eleanor and Smith-Pardo, Allan H. and Sodhi, Navjot S. and Somarriba, Eduardo J. and Sosa, Ram{\´o}n A. and Stout, Jane C. and Struebig, Matthew J. and Sung, Yik-Hei and Threlfall, Caragh G. and Tonietto, Rebecca and T{\´o}thm{\´e}r{\´e}sz, B{\´e}la and Tscharntke, Teja and Turner, Edgar C. and Tylianakis, Jason M. and Vanbergen, Adam J. and Vassilev, Kiril and Verboven, Hans A. F. and Vergara, Carlos H. and Vergara, Pablo M. and Verhulst, Jort and Walker, Tony R. and Wang, Yanping and Watling, James I. and Wells, Konstans and Williams, Christopher D. and Willig, Michael R. and Woinarski, John C. Z. and Wolf, Jan H. D. and Woodcock, Ben A. and Yu, Douglas W. and Zailsev, Andreys and Collen, Ben and Ewers, Rob M. and Mace, Georgina M. and Purves, Drew W. and Scharlemann, J{\"o}rn P. W. and Pervis, Andy}, title = {The PREDICTS database: a global database of how local terrestrial biodiversity responds to human impacts}, series = {Ecology and Evolution}, volume = {4}, journal = {Ecology and Evolution}, number = {24}, doi = {10.1002/ece3.1303}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-114425}, pages = {4701 - 4735}, year = {2014}, abstract = {Biodiversity continues to decline in the face of increasing anthropogenic pressures such as habitat destruction, exploitation, pollution and introduction of alien species. Existing global databases of species' threat status or population time series are dominated by charismatic species. The collation of datasets with broad taxonomic and biogeographic extents, and that support computation of a range of biodiversity indicators, is necessary to enable better understanding of historical declines and to project - and avert - future declines. We describe and assess a new database of more than 1.6 million samples from 78 countries representing over 28,000 species, collated from existing spatial comparisons of local-scale biodiversity exposed to different intensities and types of anthropogenic pressures, from terrestrial sites around the world. The database contains measurements taken in 208 (of 814) ecoregions, 13 (of 14) biomes, 25 (of 35) biodiversity hotspots and 16 (of 17) megadiverse countries. The database contains more than 1\% of the total number of all species described, and more than 1\% of the described species within many taxonomic groups - including flowering plants, gymnosperms, birds, mammals, reptiles, amphibians, beetles, lepidopterans and hymenopterans. The dataset, which is still being added to, is therefore already considerably larger and more representative than those used by previous quantitative models of biodiversity trends and responses. The database is being assembled as part of the PREDICTS project (Projecting Responses of Ecological Diversity In Changing Terrestrial Systems - ). We make site-level summary data available alongside this article. The full database will be publicly available in 2015.}, language = {en} } @article{BittnerBingenerCaseyDietzetal.2014, author = {Bittner, R. and Bingener-Casey, J. and Dietz, U. and Fabian, M. and Ferzli, G. S. and Fortelny, R. H. and K{\"o}ckerling, F. and Kukleta, J. and LeBlanc, K. and Lomanto, D. and Misra, M. C. and Morales-Conde, S. and Ramshaw, B. and Reinpold, W. and Rim, S. and Rohr, M. and Schrittwieser, R. and Simon, T. and Smietanski, M. and Stechemesser, B. and Timoney, M. and Chowbey, P.}, title = {Guidelines for laparoscopic treatment of ventral and incisional abdominal wall hernias (International Endohernia Society [IEHS])—Part 2}, series = {Surgical Endoscopy}, volume = {28}, journal = {Surgical Endoscopy}, number = {2}, doi = {10.1007/s00464-013-3171-5}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-121510}, pages = {353 - 379}, year = {2014}, abstract = {Guidelines are increasingly determining the decision process in day-to-day clinical work. Guidelines describe the current best possible standard in diagnostics and therapy. They should be developed by an international panel of experts, whereby alongside individual experience, above all, the results of comparative studies are decisive. According to the results of high-ranking scientific studies published in peer-reviewed journals, statements and recommendations are formulated, and these are graded strictly according to the criteria of evidence-based medicine. Guidelines can therefore be valuable in helping particularly the young surgeon in his or her day-to-day work to find the best decision for the patient when confronted with a wide and confusing range of options. However, even experienced surgeons benefit because by virtue of a heavy workload and commitment, they often find it difficult to keep up with the ever-increasing published literature. All guidelines require regular updating, usually every 3 years, in line with progress in the field. The current Guidelines focus on technique and perioperative management of laparoscopic ventral hernia repair and constitute the first comprehensive guidelines on this topic. In this issue of Surgical Endoscopy, the first part of the Guidelines is published including sections on basics, indication for surgery, perioperative management, and key points of technique. The next part (Part 2) of the Guidelines will address complications and comparisons between open and laparoscopic techniques. Part 3 will cover mesh technology, hernia prophylaxis, technique-related issues, new technologic developments, lumbar and other unusual hernias, and training/education.}, language = {en} } @article{BittnerBingenerCaseyDietzetal.2014, author = {Bittner, R. and Bingener-Casey, J. and Dietz, U. and Fabian, M. and Ferzli, G. S. and Fortelny, R. H. and K{\"o}ckerling, F. and Kukleta, J. and LeBlanc, K. and Lomanto, D. and Misra, M. C. and Bansal, V. K. and Morales-Conde, S. and Ramshaw, B. and Reinpold, W. and Rim, S. and Rohr, M. and Schrittwieser, R. and Simon, T. and Smietanski, M. and Stechemesser, B. and Timoney, M. and Chowbey, P.}, title = {Guidelines for laparoscopic treatment of ventral and incisional abdominal wall hernias (International Endohernia Society (IEHS)—Part 1}, series = {Surgical Endoscopy}, volume = {28}, journal = {Surgical Endoscopy}, doi = {10.1007/s00464-013-3170-6}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-121294}, pages = {2-29}, year = {2014}, abstract = {Guidelines are increasingly determining the decision process in day-to-day clinical work. Guidelines describe the current best possible standard in diagnostics and therapy. They should be developed by an international panel of experts, whereby alongside individual experience, above all, the results of comparative studies are decisive. According to the results of high-ranking scientific studies published in peer-reviewed journals, statements and recommendations are formulated, and these are graded strictly according to the criteria of evidence-based medicine. Guidelines can therefore be valuable in helping particularly the young surgeon in his or her day-to-day work to find the best decision for the patient when confronted with a wide and confusing range of options. However, even experienced surgeons benefit because by virtue of a heavy workload and commitment, they often find it difficult to keep up with the ever-increasing published literature. All guidelines require regular updating, usually every 3 years, in line with progress in the field. The current Guidelines focus on technique and perioperative management of laparoscopic ventral hernia repair and constitute the first comprehensive guidelines on this topic. In this issue of Surgical Endoscopy, the first part of the Guidelines is published including sections on basics, indication for surgery, perioperative management, and key points of technique. The next part (Part 2) of the Guidelines will address complications and comparisons between open and laparoscopic techniques. Part 3 will cover mesh technology, hernia prophylaxis, technique-related issues, new technologic developments, lumbar and other unusual hernias, and training/education.}, language = {en} } @article{HaringReinerLiuetal.2021, author = {Haring, Bernhard and Reiner, Alexander P. and Liu, Jungmin and Tobias, Deirdre K. and Whitsel, Eric and Berger, Jeffrey S. and Desai, Pinkal and Wassertheil-Smoller, Sylvia and LaMonte, Michael J. and Hayden, Kathleen and Bick, Alexander G. and Natarajan, Pradeep and Weinstock, Joshua S. and Nguyen, Patricia K. and Stefanick, Marcia and Simon, Michael S. and Eaton, Charles and Kooperberg, Charles and Manson, JoAnn E.}, title = {Healthy Lifestyle and Clonal Hematopoiesis of Indeterminate Potential. Results from the Women's Health Initiative}, series = {Journal of the American Heart Association}, volume = {10}, journal = {Journal of the American Heart Association}, number = {5}, doi = {10.1161/JAHA.120.018789}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-236146}, year = {2021}, abstract = {Background Presence of clonal hematopoiesis of indeterminate potential (CHIP) is associated with a higher risk of atherosclerotic cardiovascular disease, cancer, and mortality. The relationship between a healthy lifestyle and CHIP is unknown. Methods and Results This analysis included 8709 postmenopausal women (mean age, 66.5 years) enrolled in the WHI (Women's Health Initiative), free of cancer or cardiovascular disease, with deep-coverage whole genome sequencing data available. Information on lifestyle factors (body mass index, smoking, physical activity, and diet quality) was obtained, and a healthy lifestyle score was created on the basis of healthy criteria met (0 point [least healthy] to 4 points [most healthy]). CHIP was derived on the basis of a prespecified list of leukemogenic driver mutations. The prevalence of CHIP was 8.6\%. A higher healthy lifestyle score was not associated with CHIP (multivariable-adjusted odds ratio [OR] [95\% CI], 0.99 [0.80-1.23] and 1.13 [0.93-1.37]) for the upper (3 or 4 points) and middle category (2 points), respectively, versus referent (0 or 1 point). Across score components, a normal and overweight body mass index compared with obese was significantly associated with a lower odds for CHIP (OR, 0.71 [95\% CI, 0.57-0.88] and 0.83 [95\% CI, 0.68-1.01], respectively; P-trend 0.0015). Having never smoked compared with being a current smoker tended to be associated with lower odds for CHIP. Conclusions A healthy lifestyle, based on a composite score, was not related to CHIP among postmenopausal women. However, across individual lifestyle factors, having a normal body mass index was strongly associated with a lower prevalence of CHIP. These findings support the idea that certain healthy lifestyle factors are associated with a lower frequency of CHIP.}, language = {en} } @article{BittnerBingenerCaseyDietzetal.2014, author = {Bittner, R. and Bingener-Casey, J. and Dietz, U. and Fabian, M. and Ferzli, G. and Fortelny, R. and K{\"o}ckerling, F. and Kukleta, J. and LeBlanc, K. and Lomanto, D. and Misra, M. and Morales-Conde, S. and Ramshaw, B. and Reinpold, W. and Rim, S. and Rohr, M. and Schrittwieser, R. and Simon, T. and Smietanski, M. and Stechemesser, B. and Timoney, M. and Chowbey, P.}, title = {Guidelines for laparoscopic treatment of ventral and incisional abdominal wall hernias (International Endohernia Society [IEHS])—Part III}, series = {Surgical Endoscopy}, volume = {28}, journal = {Surgical Endoscopy}, number = {2}, doi = {10.1007/s00464-013-3172-4}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-121289}, pages = {380-404}, year = {2014}, abstract = {Guidelines are increasingly determining the decision process in day-to-day clinical work. Guidelines describe the current best possible standard in diagnostics and therapy. They should be developed by an international panel of experts, whereby alongside individual experience, above all, the results of comparative studies are decisive. According to the results of high-ranking scientific studies published in peer-reviewed journals, statements and recommendations are formulated, and these are graded strictly according to the criteria of evidence-based medicine. Guidelines can therefore be valuable in helping particularly the young surgeon in his or her day-to-day work to find the best decision for the patient when confronted with a wide and confusing range of options. However, even experienced surgeons benefit because by virtue of a heavy workload and commitment, they often find it difficult to keep up with the ever-increasing published literature. All guidelines require regular updating, usually every 3 years, in line with progress in the field. The current Guidelines focus on technique and perioperative management of laparoscopic ventral hernia repair and constitute the first comprehensive guidelines on this topic. In this issue of Surgical Endoscopy, the first part of the Guidelines is published including sections on basics, indication for surgery, perioperative management, and key points of technique. The next part (Part 2) of the Guidelines will address complications and comparisons between open and laparoscopic techniques. Part 3 will cover mesh technology, hernia prophylaxis, technique-related issues, new technologic developments, lumbar and other unusual hernias, and training/education.}, language = {en} } @article{AntoniouKuchenbaeckerSoucyetal.2012, author = {Antoniou, Antonis C. and Kuchenbaecker, Karoline B. and Soucy, Penny and Beesley, Jonathan and Chen, Xiaoqing and McGuffog, Lesley and Lee, Andrew and Barrowdale, Daniel and Healey, Sue and Sinilnikova, Olga M. and Caligo, Maria A. and Loman, Niklas and Harbst, Katja and Lindblom, Annika and Arver, Brita and Rosenquist, Richard and Karlsson, Per and Nathanson, Kate and Domchek, Susan and Rebbeck, Tim and Jakubowska, Anna and Lubinski, Jan and Jaworska, Katarzyna and Durda, Katarzyna and Zlowowcka-Perłowska, Elżbieta and Osorio, Ana and Dur{\´a}n, Mercedes and Andr{\´e}s, Raquel and Ben{\´i}tez, Javier and Hamann, Ute and Hogervorst, Frans B. and van Os, Theo A. and Verhoef, Senno and Meijers-Heijboer, Hanne E. J. and Wijnen, Juul and Garcia, Encarna B. G{\´o}mez and Ligtenberg, Marjolijn J. and Kriege, Mieke and Coll{\´e}e, Margriet and Ausems, Margreet G. E. M. and Oosterwijk, Jan C. and Peock, Susan and Frost, Debra and Ellis, Steve D. and Platte, Radka and Fineberg, Elena and Evans, D. Gareth and Lalloo, Fiona and Jacobs, Chris and Eeles, Ros and Adlard, Julian and Davidson, Rosemarie and Cole, Trevor and Cook, Jackie and Paterson, Joan and Douglas, Fiona and Brewer, Carole and Hodgson, Shirley and Morrison, Patrick J. and Walker, Lisa and Rogers, Mark T. and Donaldson, Alan and Dorkins, Huw and Godwin, Andrew K. and Bove, Betsy and Stoppa-Lyonnet, Dominique and Houdayer, Claude and Buecher, Bruno and de Pauw, Antoine and Mazoyer, Sylvie and Calender, Alain and L{\´e}on{\´e}, M{\´e}lanie and Bressac-de Paillerets, Brigitte and Caron, Olivier and Sobol, Hagay and Frenay, Marc and Prieur, Fabienne and Ferrer, Sandra Fert and Mortemousque, Isabelle and Buys, Saundra and Daly, Mary and Miron, Alexander and Terry, Mary Beth and Hopper, John L. and John, Esther M. and Southey, Melissa and Goldgar, David and Singer, Christian F. and Fink-Retter, Anneliese and Muy-Kheng, Tea and Geschwantler Kaulich, Daphne and Hansen, Thomas V. O. and Nielsen, Finn C. and Barkardottir, Rosa B. and Gaudet, Mia and Kirchhoff, Tomas and Joseph, Vijai and Dutra-Clarke, Ana and Offit, Kenneth and Piedmonte, Marion and Kirk, Judy and Cohn, David and Hurteau, Jean and Byron, John and Fiorica, James and Toland, Amanda E. and Montagna, Marco and Oliani, Cristina and Imyanitov, Evgeny and Isaacs, Claudine and Tihomirova, Laima and Blanco, Ignacio and Lazaro, Conxi and Teul{\´e}, Alex and Del Valle, J. and Gayther, Simon A. and Odunsi, Kunle and Gross, Jenny and Karlan, Beth Y. and Olah, Edith and Teo, Soo-Hwang and Ganz, Patricia A. and Beattie, Mary S. and Dorfling, Cecelia M. and Jansen van Rensburg, Elizabeth and Diez, Orland and Kwong, Ava and Schmutzler, Rita K. and Wappenschmidt, Barbara and Engel, Christoph and Meindl, Alfons and Ditsch, Nina and Arnold, Norbert and Heidemann, Simone and Niederacher, Dieter and Preisler-Adams, Sabine and Gadzicki, Dorothea and Varon-Mateeva, Raymonda and Deissler, Helmut and Gehrig, Andrea and Sutter, Christian and Kast, Karin and Fiebig, Britta and Sch{\"a}fer, Dieter and Caldes, Trinidad and de la Hoya, Miguel and Nevanlinna, Heli and Muranen, Taru A. and Lesp{\´e}rance, Bernard and Spurdle, Amanda B. and Neuhausen, Susan L. and Ding, Yuan C. and Wang, Xianshu and Fredericksen, Zachary and Pankratz, Vernon S. and Lindor, Noralane M. and Peterlongo, Paulo and Manoukian, Siranoush and Peissel, Bernard and Zaffaroni, Daniela and Bonanni, Bernardo and Bernard, Loris and Dolcetti, Riccardo and Papi, Laura and Ottini, Laura and Radice, Paolo and Greene, Mark H. and Loud, Jennifer T. and Andrulis, Irene L. and Ozcelik, Hilmi and Mulligan, Anna Marie and Glendon, Gord and Thomassen, Mads and Gerdes, Anne-Marie and Jensen, Uffe B. and Skytte, Anne-Bine and Kruse, Torben A. and Chenevix-Trench, Georgia and Couch, Fergus J. and Simard, Jacques and Easton, Douglas F.}, title = {Common variants at 12p11, 12q24, 9p21, 9q31.2 and in ZNF365 are associated with breast cancer risk for BRCA1 and/or BRCA2 mutation carriers}, series = {Breast Cancer Research}, volume = {14}, journal = {Breast Cancer Research}, number = {R33}, organization = {CIMBA; SWE-BRCA; HEBON; EMBRACE; GEMO Study Collaborators; kConFab Investigators}, doi = {10.1186/bcr3121}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-130449}, year = {2012}, abstract = {Introduction: Several common alleles have been shown to be associated with breast and/or ovarian cancer risk for BRCA1 and BRCA2 mutation carriers. Recent genome-wide association studies of breast cancer have identified eight additional breast cancer susceptibility loci: rs1011970 (9p21, CDKN2A/B), rs10995190 (ZNF365), rs704010 (ZMIZ1), rs2380205 (10p15), rs614367 (11q13), rs1292011 (12q24), rs10771399 (12p11 near PTHLH) and rs865686 (9q31.2). Methods: To evaluate whether these single nucleotide polymorphisms (SNPs) are associated with breast cancer risk for BRCA1 and BRCA2 carriers, we genotyped these SNPs in 12,599 BRCA1 and 7,132 BRCA2 mutation carriers and analysed the associations with breast cancer risk within a retrospective likelihood framework. Results: Only SNP rs10771399 near PTHLH was associated with breast cancer risk for BRCA1 mutation carriers (per-allele hazard ratio (HR) = 0.87, 95\% CI: 0.81 to 0.94, P-trend = 3 x 10\(^{-4}\)). The association was restricted to mutations proven or predicted to lead to absence of protein expression (HR = 0.82, 95\% CI: 0.74 to 0.90, P-trend = 3.1 x 10\(^{-5}\), P-difference = 0.03). Four SNPs were associated with the risk of breast cancer for BRCA2 mutation carriers: rs10995190, P-trend = 0.015; rs1011970, P-trend = 0.048; rs865686, 2df P = 0.007; rs1292011 2df P = 0.03. rs10771399 (PTHLH) was predominantly associated with estrogen receptor (ER)-negative breast cancer for BRCA1 mutation carriers (HR = 0.81, 95\% CI: 0.74 to 0.90, P-trend = 4 x 10\(^{-5}\)) and there was marginal evidence of association with ER- negative breast cancer for BRCA2 mutation carriers (HR = 0.78, 95\% CI: 0.62 to 1.00, P-trend = 0.049). Conclusions: The present findings, in combination with previously identified modifiers of risk, will ultimately lead to more accurate risk prediction and an improved understanding of the disease etiology in BRCA1 and BRCA2 mutation carriers.}, language = {en} } @article{MitchellMacarthurGanetal.2014, author = {Mitchell, Anna L. and Macarthur, Katie D. R. and Gan, Earn H. and Baggott, Lucy E. and Wolff, Anette S. B. and Skinningsrud, Beate and Platt, Hazel and Short, Andrea and Lobell, Anna and Kampe, Olle and Bensing, Sophie and Betterle, Corrado and Kasperlik-Zaluska, Anna and Zurawek, Magdalena and Fichna, Marta and Kockum, Ingrid and Eriksson, Gabriel Nordling and Ekwall, Olov and Wahlberg, Jeanette and Dahlqvist, Per and Hulting, Anna-Lena and Penna-Martinez, Marissa and Meyer, Gesine and Kahles, Heinrich and Badenhoop, Klaus and Hahner, Stephanie and Quinkler, Marcus and Falorni, Alberto and Phipps-Green, Amanda and Merriman, Tony R. and Ollier, William and Cordell, Heather J. and Undlien, Dag and Czarnocka, Barbara and Husebye, Eystein and Pearce, Simon H. S.}, title = {Association of Autoimmune Addison's Disease with Alleles of STAT4 and GATA3 in European Cohorts}, series = {PLOS ONE}, volume = {9}, journal = {PLOS ONE}, number = {3}, doi = {10.1371/journal.pone.0088991}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-117105}, pages = {e88991}, year = {2014}, abstract = {Background: Gene variants known to contribute to Autoimmune Addison's disease (AAD) susceptibility include those at the MHC, MICA, CIITA, CTLA4, PTPN22, CYP27B1, NLRP-1 and CD274 loci. The majority of the genetic component to disease susceptibility has yet to be accounted for. Aim: To investigate the role of 19 candidate genes in AAD susceptibility in six European case-control cohorts. Methods: A sequential association study design was employed with genotyping using Sequenom iPlex technology. In phase one, 85 SNPs in 19 genes were genotyped in UK and Norwegian AAD cohorts (691 AAD, 715 controls). In phase two, 21 SNPs in 11 genes were genotyped in German, Swedish, Italian and Polish cohorts (1264 AAD, 1221 controls). In phase three, to explore association of GATA3 polymorphisms with AAD and to determine if this association extended to other autoimmune conditions, 15 SNPs in GATA3 were studied in UK and Norwegian AAD cohorts, 1195 type 1 diabetes patients from Norway, 650 rheumatoid arthritis patients from New Zealand and in 283 UK Graves' disease patients. Meta-analysis was used to compare genotype frequencies between the participating centres, allowing for heterogeneity. Results: We report significant association with alleles of two STAT4 markers in AAD cohorts (rs4274624: P = 0.00016; rs10931481: P = 0.0007). In addition, nominal association of AAD with alleles at GATA3 was found in 3 patient cohorts and supported by meta-analysis. Association of AAD with CYP27B1 alleles was also confirmed, which replicates previous published data. Finally, nominal association was found at SNPs in both the NF-kappa B1 and IL23A genes in the UK and Italian cohorts respectively. Conclusions: Variants in the STAT4 gene, previously associated with other autoimmune conditions, confer susceptibility to AAD. Additionally, we report association of GATA3 variants with AAD: this adds to the recent report of association of GATA3 variants with rheumatoid arthritis.}, language = {en} } @article{KleijnWinfreeBartomeusetal.2015, author = {Kleijn, David and Winfree, Rachael and Bartomeus, Ignasi and Carvalheiro, Lu{\´i}sa G. and Henry, Mickael and Isaacs, Rufus and Klein, Alexandra-Maria and Kremen, Claire and M'Gonigle, Leithen K. and Rader, Romina and Ricketts, Taylor H. and Williams, Neal M. and Adamson, Nancy Lee and Ascher, John S. and B{\´a}ldi, Andr{\´a}s and Bat{\´a}ry, P{\´e}ter and Benjamin, Faye and Biesmeijer, Jacobus C. and Blitzer, Eleanor J. and Bommarco, Riccardo and Brand, Mariette R. and Bretagnolle, Vincent and Button, Lindsey and Cariveau, Daniel P. and Chifflet, R{\´e}my and Colville, Jonathan F. and Danforth, Bryan N. and Elle, Elizabeth and Garratt, Michael P. D. and Herzog, Felix and Holzschuh, Andrea and Howlett, Brad G. and Jauker, Frank and Jha, Shalene and Knop, Eva and Krewenka, Kristin M. and Le F{\´e}on, Violette and Mandelik, Yael and May, Emily A. and Park, Mia G. and Pisanty, Gideon and Reemer, Menno and Riedinger, Verena and Rollin, Orianne and Rundl{\"o}f, Maj and Sardi{\~n}as, Hillary S. and Scheper, Jeroen and Sciligo, Amber R. and Smith, Henrik G. and Steffan-Dewenter, Ingolf and Thorp, Robbin and Tscharntke, Teja and Verhulst, Jort and Viana, Blandina F. and Vaissi{\`e}re, Bernard E. and Veldtman, Ruan and Ward, Kimiora L. and Westphal, Catrin and Potts, Simon G.}, title = {Delivery of crop pollination services is an insufficient argument for wild pollinator conservation}, series = {Nature Communications}, volume = {6}, journal = {Nature Communications}, number = {7414}, doi = {10.1038/ncomms8414}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-151879}, year = {2015}, abstract = {There is compelling evidence that more diverse ecosystems deliver greater benefits to people, and these ecosystem services have become a key argument for biodiversity conservation. However, it is unclear how much biodiversity is needed to deliver ecosystem services in a cost- effective way. Here we show that, while the contribution of wild bees to crop production is significant, service delivery is restricted to a limited subset of all known bee species. Across crops, years and biogeographical regions, crop-visiting wild bee communities are dominated by a small number of common species, and threatened species are rarely observed on crops. Dominant crop pollinators persist under agricultural expansion and many are easily enhanced by simple conservation measures, suggesting that cost- effective management strategies to promote crop pollination should target a different set of species than management strategies to promote threatened bees. Conserving the biological diversity of bees therefore requires more than just ecosystem-service-based arguments.}, language = {en} } @article{WurmbSchorscherJusticeetal.2018, author = {Wurmb, T and Schorscher, N and Justice, P and Dietz, S and Schua, R and Jarausch, T and Kinstle, U and Greiner, J and M{\"o}ldner, G and M{\"u}ller, J and Kraus, M and Simon, S and Wagenh{\"a}user, U and Hemm, J and Roewer, N and Helm, M}, title = {Structured analysis, evaluation and report of the emergency response to a terrorist attack in Wuerzburg, Germany using a new template of standardised quality indicators}, series = {Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine}, volume = {26}, journal = {Scandinavian Journal of Trauma, Resuscitation and Emergency Medicine}, number = {87}, doi = {10.1186/s13049-018-0555-5}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-177054}, year = {2018}, abstract = {Background: Until now there has been a reported lack of systematic reports and scientific evaluations of rescue missions during terror attacks. This however is urgently required in order to improve the performance of emergency medical services and to be able to compare different missions with each other. Aim of the presented work was to report the systematic evaluation and the lessons learned from the response to a terror attack that happened in Wuerzburg, Germany in 2016. Methods: A team of 14 experts developed a template of quality indicators and operational characteristics, which allow for the description, assessment and comparison of civil emergency rescue missions during mass killing incidents. The entire systematic evaluation process consisted of three main steps. The first step was the systematic data collection according to the quality indicators and operational characteristics. Second was the systematic stratification and assessment of the data. The last step was the prioritisation of the identified weaknesses and the definition of the lessons learned. Results: Five important "lessons learned" have been defined. First of all, a comprehensive concept for rescue missions during terror attacks is essential. Furthermore, the establishment of a defined high priority communication infrastructure between the different dispatch centres ("red phone") is vital. The goal is to secure the continuity of information between a few well-defined individuals. Thirdly, the organization of the incident scene needs to be commonly decided and communicated between police, medical services and fire services during the mission. A successful mission tactic requires continuous flux of reports to the on-site command post. Therefore, a predefined and common communication infrastructure for all operational forces is a crucial point. Finally, all strategies need to be extensively trained before the real life scenario hits. Conclusion: According to a systematic evaluation, we defined the lessons learned from a terror attack in 2016. Further systematic reports and academic work surrounding life threatening rescue missions and mass killing incidents are needed in order to ultimately improve such mission outcomes. In the future, a close international collaboration might help to find the best database to report and evaluate major incidents but also mass killing events.}, language = {en} } @article{DoerkPeterlongoMannermaaetal.2019, author = {D{\"o}rk, Thilo and Peterlongo, Peter and Mannermaa, Arto and Bolla, Manjeet K. and Wang, Qin and Dennis, Joe and Ahearn, Thomas and Andrulis, Irene L. and Anton-Culver, Hoda and Arndt, Volker and Aronson, Kristan J. and Augustinsson, Annelie and Beane Freeman, Laura E. and Beckmann, Matthias W. and Beeghly-Fadiel, Alicia and Behrens, Sabine and Bermisheva, Marina and Blomqvist, Carl and Bogdanova, Natalia V. and Bojesen, Stig E. and Brauch, Hiltrud and Brenner, Hermann and Burwinkel, Barbara and Canzian, Federico and Chan, Tsun L. and Chang-Claude, Jenny and Chanock, Stephen J. and Choi, Ji-Yeob and Christiansen, Hans and Clarke, Christine L. and Couch, Fergus J. and Czene, Kamila and Daly, Mary B. and dos-Santos-Silva, Isabel and Dwek, Miriam and Eccles, Diana M. and Ekici, Arif B. and Eriksson, Mikael and Evans, D. Gareth and Fasching, Peter A. and Figueroa, Jonine and Flyger, Henrik and Fritschi, Lin and Gabrielson, Marike and Gago-Dominguez, Manuela and Gao, Chi and Gapstur, Susan M. and Garc{\´i}a-Closas, Montserrat and Garc{\´i}a-S{\´a}enz, Jos{\´e} A. and Gaudet, Mia M. and Giles, Graham G. and Goldberg, Mark S. and Goldgar, David E. and Guen{\´e}l, Pascal and Haeberle, Lothar and Haimann, Christopher A. and H{\aa}kansson, Niclas and Hall, Per and Hamann, Ute and Hartman, Mikael and Hauke, Jan and Hein, Alexander and Hillemanns, Peter and Hogervorst, Frans B. L. and Hooning, Maartje J. and Hopper, John L. and Howell, Tony and Huo, Dezheng and Ito, Hidemi and Iwasaki, Motoki and Jakubowska, Anna and Janni, Wolfgang and John, Esther M. and Jung, Audrey and Kaaks, Rudolf and Kang, Daehee and Kapoor, Pooja Middha and Khusnutdinova, Elza and Kim, Sung-Won and Kitahara, Cari M. and Koutros, Stella and Kraft, Peter and Kristensen, Vessela N. and Kwong, Ava and Lambrechts, Diether and Le Marchand, Loic and Li, Jingmei and Lindstr{\"o}m, Sara and Linet, Martha and Lo, Wing-Yee and Long, Jirong and Lophatananon, Artitaya and Lubiński, Jan and Manoochehri, Mehdi and Manoukian, Siranoush and Margolin, Sara and Martinez, Elena and Matsuo, Keitaro and Mavroudis, Dimitris and Meindl, Alfons and Menon, Usha and Milne, Roger L. and Mohd Taib, Nur Aishah and Muir, Kenneth and Mulligan, Anna Marie and Neuhausen, Susan L. and Nevanlinna, Heli and Neven, Patrick and Newman, William G. and Offit, Kenneth and Olopade, Olufunmilayo I. and Olshan, Andrew F. and Olson, Janet E. and Olsson, H{\aa}kan and Park, Sue K. and Park-Simon, Tjoung-Won and Peto, Julian and Plaseska-Karanfilska, Dijana and Pohl-Rescigno, Esther and Presneau, Nadege and Rack, Brigitte and Radice, Paolo and Rashid, Muhammad U. and Rennert, Gad and Rennert, Hedy S. and Romero, Atocha and Ruebner, Matthias and Saloustros, Emmanouil and Schmidt, Marjanka K. and Schmutzler, Rita K. and Schneider, Michael O. and Schoemaker, Minouk J. and Scott, Christopher and Shen, Chen-Yang and Shu, Xiao-Ou and Simard, Jaques and Slager, Susan and Smichkoska, Snezhana and Southey, Melissa C. and Spinelli, John J. and Stone, Jennifer and Surowy, Harald and Swerdlow, Anthony J. and Tamimi, Rulla M. and Tapper, William J. and Teo, Soo H. and Terry, Mary Beth and Toland, Amanda E. and Tollenaar, Rob A. E. M. and Torres, Diana and Torres-Mej{\´i}a, Gabriela and Troester, Melissa A. and Truong, Th{\´e}r{\`e}se and Tsugane, Shoichiro and Untch, Michael and Vachon, Celine M. and van den Ouweland, Ans M. W. and van Veen, Elke M. and Vijai, Joseph and Wendt, Camilla and Wolk, Alicja and Yu, Jyh-Cherng and Zheng, Wei and Ziogas, Argyrios and Ziv, Elad and Dunnig, Alison and Pharaoh, Paul D. P. and Schindler, Detlev and Devilee, Peter and Easton, Douglas F.}, title = {Two truncating variants in FANCC and breast cancer risk}, series = {Scientific Reports}, volume = {9}, journal = {Scientific Reports}, organization = {ABCTB Investigators, NBCS Collaborators}, doi = {10.1038/s41598-019-48804-y}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-222838}, year = {2019}, abstract = {Fanconi anemia (FA) is a genetically heterogeneous disorder with 22 disease-causing genes reported to date. In some FA genes, monoallelic mutations have been found to be associated with breast cancer risk, while the risk associations of others remain unknown. The gene for FA type C, FANCC, has been proposed as a breast cancer susceptibility gene based on epidemiological and sequencing studies. We used the Oncoarray project to genotype two truncating FANCC variants (p.R185X and p.R548X) in 64,760 breast cancer cases and 49,793 controls of European descent. FANCC mutations were observed in 25 cases (14 with p.R185X, 11 with p.R548X) and 26 controls (18 with p.R185X, 8 with p.R548X). There was no evidence of an association with the risk of breast cancer, neither overall (odds ratio 0.77, 95\%CI 0.44-1.33, p = 0.4) nor by histology, hormone receptor status, age or family history. We conclude that the breast cancer risk association of these two FANCC variants, if any, is much smaller than for BRCA1, BRCA2 or PALB2 mutations. If this applies to all truncating variants in FANCC it would suggest there are differences between FA genes in their roles on breast cancer risk and demonstrates the merit of large consortia for clarifying risk associations of rare variants.}, language = {en} } @article{MemmelSisarioZoelleretal.2017, author = {Memmel, Simon and Sisario, Dmitri and Z{\"o}ller, Caren and Fiedler, Vanessa and Katzer, Astrid and Heiden, Robin and Becker, Nicholas and Eing, Lorenz and Ferreira, F{\´a}bio L.R. and Zimmermann, Heiko and Sauer, Markus and Flentje, Michael and Sukhorukov, Vladimir L. and Djuzenova, Cholpon S.}, title = {Migration pattern, actin cytoskeleton organization and response to PI3K-, mTOR-, and Hsp90-inhibition of glioblastoma cells with different invasive capacities}, series = {Oncotarget}, volume = {8}, journal = {Oncotarget}, number = {28}, doi = {10.18632/oncotarget.16847}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-170719}, pages = {45298-45310}, year = {2017}, abstract = {High invasiveness and resistance to chemo- and radiotherapy of glioblastoma multiforme (GBM) make it the most lethal brain tumor. Therefore, new treatment strategies for preventing migration and invasion of GBM cells are needed. Using two different migration assays, Western blotting, conventional and super-resolution (dSTORM) fluorescence microscopy we examine the effects of the dual PI3K/mTOR-inhibitor PI-103 alone and in combination with the Hsp90 inhibitor NVP-AUY922 and/or irradiation on the migration, expression of marker proteins, focal adhesions and F-actin cytoskeleton in two GBM cell lines (DK-MG and SNB19) markedly differing in their invasive capacity. Both lines were found to be strikingly different in morphology and migration behavior. The less invasive DK-MG cells maintained a polarized morphology and migrated in a directionally persistent manner, whereas the highly invasive SNB19 cells showed a multipolar morphology and migrated randomly. Interestingly, a single dose of 2 Gy accelerated wound closure in both cell lines without affecting their migration measured by single-cell tracking. PI-103 inhibited migration of DK-MG (p53 wt, PTEN wt) but not of SNB19 (p53 mut, PTEN mut) cells probably due to aberrant reactivation of the PI3K pathway in SNB19 cells treated with PI-103. In contrast, NVP-AUY922 exerted strong anti-migratory effects in both cell lines. Inhibition of cell migration was associated with massive morphological changes and reorganization of the actin cytoskeleton. Our results showed a cell line-specific response to PI3K/mTOR inhibition in terms of GBM cell motility. We conclude that anti-migratory agents warrant further preclinical investigation as potential therapeutics for treatment of GBM.}, language = {en} } @article{HarterBernatzScholzetal.2015, author = {Harter, Patrick N. and Bernatz, Simon and Scholz, Alexander and Zeiner, Pia S. and Zinke, Jenny and Kiyose, Makoto and Blasel, Stella and Beschorner, Rudi and Senft, Christian and Bender, Benjamin and Ronellenfitsch, Michael W. and Wikman, Harriet and Glatzel, Markus and Meinhardt, Matthias and Juratli, Tareq A. and Steinbach, Joachim P. and Plate, Karl H. and Wischhusen, J{\"o}rg and Weide, Benjamin and Mittelbronn, Michel}, title = {Distribution and prognostic relevance of tumor-infiltrating lymphocytes (TILs) and PD-1/PD-L1 immune checkpoints in human brain metastases}, series = {Oncotarget}, volume = {6}, journal = {Oncotarget}, number = {38}, doi = {10.18632/oncotarget.5696}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-137107}, pages = {40836 -- 40849}, year = {2015}, abstract = {The activation of immune cells by targeting checkpoint inhibitors showed promising results with increased patient survival in distinct primary cancers. Since only limited data exist for human brain metastases, we aimed at characterizing tumor infiltrating lymphocytes (TILs) and expression of immune checkpoints in the respective tumors. Two brain metastases cohorts, a mixed entity cohort (n = 252) and a breast carcinoma validation cohort (n = 96) were analyzed for CD3+, CD8+, FOXP3+, PD-1+ lymphocytes and PD-L1+ tumor cells by immunohistochemistry. Analyses for association with clinico-epidemiological and neuroradiological parameters such as patient survival or tumor size were performed. TILs infiltrated brain metastases in three different patterns (stromal, peritumoral, diffuse). While carcinomas often show a strong stromal infiltration, TILs in melanomas often diffusely infiltrate the tumors. Highest levels of CD3+ and CD8+ lymphocytes were seen in renal cell carcinomas (RCC) and strongest PD-1 levels on RCCs and melanomas. High amounts of TILs, high ratios of PD-1+/CD8+ cells and high levels of PD-L1 were negatively correlated with brain metastases size, indicating that in smaller brain metastases CD8+ immune response might get blocked. PD-L1 expression strongly correlated with TILs and FOXP3 expression. No significant association of patient survival with TILs was observed, while high levels of PD-L1 showed a strong trend towards better survival in melanoma brain metastases (Log-Rank p = 0.0537). In summary, melanomas and RCCs seem to be the most immunogenic entities. Differences in immunotherapeutic response between tumor entities regarding brain metastases might be attributable to this finding and need further investigation in larger patient cohorts.}, language = {en} } @article{HuserRohwedderApostolopoulouetal.2012, author = {Huser, Annina and Rohwedder, Astrid and Apostolopoulou, Anthi A. and Widmann, Annekathrin and Pfitzenmaier, Johanna E. and Maiolo, Elena M. and Selcho, Mareike and Pauls, Dennis and von Essen, Alina and Gupta, Tript and Sprecher, Simon G. and Birman, Serge and Riemensperger, Thomas and Stocker, Reinhard F. and Thum, Andreas S.}, title = {The Serotonergic Central Nervous System of the Drosophila Larva: Anatomy and Behavioral Function}, series = {PLoS One}, volume = {7}, journal = {PLoS One}, number = {10}, doi = {10.1371/journal.pone.0047518}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-130437}, pages = {e47518}, year = {2012}, abstract = {The Drosophila larva has turned into a particularly simple model system for studying the neuronal basis of innate behaviors and higher brain functions. Neuronal networks involved in olfaction, gustation, vision and learning and memory have been described during the last decade, often up to the single-cell level. Thus, most of these sensory networks are substantially defined, from the sensory level up to third-order neurons. This is especially true for the olfactory system of the larva. Given the wealth of genetic tools in Drosophila it is now possible to address the question how modulatory systems interfere with sensory systems and affect learning and memory. Here we focus on the serotonergic system that was shown to be involved in mammalian and insect sensory perception as well as learning and memory. Larval studies suggested that the serotonergic system is involved in the modulation of olfaction, feeding, vision and heart rate regulation. In a dual anatomical and behavioral approach we describe the basic anatomy of the larval serotonergic system, down to the single-cell level. In parallel, by expressing apoptosis-inducing genes during embryonic and larval development, we ablate most of the serotonergic neurons within the larval central nervous system. When testing these animals for naive odor, sugar, salt and light perception, no profound phenotype was detectable; even appetitive and aversive learning was normal. Our results provide the first comprehensive description of the neuronal network of the larval serotonergic system. Moreover, they suggest that serotonin per se is not necessary for any of the behaviors tested. However, our data do not exclude that this system may modulate or fine-tune a wide set of behaviors, similar to its reported function in other insect species or in mammals. Based on our observations and the availability of a wide variety of genetic tools, this issue can now be addressed.}, language = {en} } @article{DjuzenovaMemmelSukhorukovetal.2014, author = {Djuzenova, Cholpon S. and Memmel, Simon and Sukhorukov, Vladimir L. and H{\"o}ring, Marcus and Westerling, Katherine and Fiedler, Vanessa and Katzer, Astrid and Krohne, Georg and Flentje, Michael}, title = {Cell Surface Area and Membrane Folding in Glioblastoma Cell Lines Differing in PTEN and p53 Status}, doi = {10.1371/journal.pone.0087052}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-111322}, year = {2014}, abstract = {Glioblastoma multiforme (GBM) is characterized by rapid growth, invasion and resistance to chemo-/radiotherapy. The complex cell surface morphology with abundant membrane folds, microvilli, filopodia and other membrane extensions is believed to contribute to the highly invasive behavior and therapy resistance of GBM cells. The present study addresses the mechanisms leading to the excessive cell membrane area in five GBM lines differing in mutational status for PTEN and p53. In addition to scanning electron microscopy (SEM), the membrane area and folding were quantified by dielectric measurements of membrane capacitance using the single-cell electrorotation (ROT) technique. The osmotic stability and volume regulation of GBM cells were analyzed by video microscopy. The expression of PTEN, p53, mTOR and several other marker proteins involved in cell growth and membrane synthesis were examined by Western blotting. The combined SEM, ROT and osmotic data provided independent lines of evidence for a large variability in membrane area and folding among tested GBM lines. Thus, DK-MG cells (wild type p53 and wild type PTEN) exhibited the lowest degree of membrane folding, probed by the area-specific capacitance Cm = 1.9 µF/cm2. In contrast, cell lines carrying mutations in both p53 and PTEN (U373-MG and SNB19) showed the highest Cm values of 3.7-4.0 µF/cm2, which corroborate well with their heavily villated cell surface revealed by SEM. Since PTEN and p53 are well-known inhibitors of mTOR, the increased membrane area/folding in mutant GBM lines may be related to the enhanced protein and lipid synthesis due to a deregulation of the mTOR-dependent downstream signaling pathway. Given that membrane folds and extensions are implicated in tumor cell motility and metastasis, the dielectric approach presented here provides a rapid and simple tool for screening the biophysical cell properties in studies on targeting chemo- or radiotherapeutically the migration and invasion of GBM and other tumor types.}, language = {en} } @article{ChenWernerLapaetal.2018, author = {Chen, Xinyu and Werner, Rudolf A. and Lapa, Constantin and Nose, Naoko and Hirano, Mitsuru and Javadi, Mehrbod S. and Robinson, Simon and Higuchi, Takahiro}, title = {Subcellular storage and release mode of the novel \(^{18}\)F-labeled sympathetic nerve PET tracer LMI1195}, series = {EJNMMI Research}, volume = {8}, journal = {EJNMMI Research}, number = {12}, issn = {2191-219X}, doi = {10.1186/s13550-018-0365-9}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-167081}, year = {2018}, abstract = {Background: \(^{18}\)F-N-[3-bromo-4-(3-fluoro-propoxy)-benzyl]-guanidine (\(^{18}\)F-LMI1195) is a new class of PET tracer designed for sympathetic nervous imaging of the heart. The favorable image quality with high and specific neural uptake has been previously demonstrated in animals and humans, but intracellular behavior is not yet fully understood. The aim of the present study is to verify whether it is taken up in storage vesicles and released in company with vesicle turnover. Results: Both vesicle-rich (PC12) and vesicle-poor (SK-N-SH) norepinephrine-expressing cell lines were used for in vitro tracer uptake studies. After 2 h of \(^{18}\)F-LMI1195 preloading into both cell lines, effects of stimulants for storage vesicle turnover (high concentration KCl (100 mM) or reserpine treatment) were measured at 10, 20, and 30 min. \(^{131}\)I-meta-iodobenzylguanidine (\(^{131}\)I-MIBG) served as a reference. Both high concentration KCl and reserpine enhanced \(^{18}\)F-LMI1195 washout from PC12 cells, while tracer retention remained stable in the SK-N-SH cells. After 30 min of treatment, 18F-LMI1195 releasing index (percentage of tracer released from cells) from vesicle-rich PC12 cells achieved significant differences compared to cells without treatment condition. In contrast, such effect could not be observed using vesicle-poor SK-N-SH cell lines. Similar tracer kinetics after KCl or reserpine treatment were also observed using 131I-MIBG. In case of KCl exposure, Ca\(^{2+}\)-free buffer with the calcium chelator, ethylenediaminetetracetic acid (EDTA), could suppress the tracer washout from PC12 cells. This finding is consistent with the tracer release being mediated by Ca\(^{2+}\) influx resulting from membrane depolarization. Conclusions: Analogous to \(^{131}\)I-MIBG, the current in vitro tracer uptake study confirmed that \(^{131}\)F-LMI1195 is also stored in vesicles in PC12 cells and released along with vesicle turnover. Understanding the basic kinetics of \(^{18}\)FLMI1195 at a subcellular level is important for the design of clinical imaging protocols and imaging interpretation.}, subject = {Positronen-Emissions-Tomografie}, language = {en} } @article{PostemaHoogmanAmbrosinoetal.2021, author = {Postema, Merel C. and Hoogman, Martine and Ambrosino, Sara and Asherson, Philip and Banaschewski, Tobias and Bandeira, Cibele E. and Baranov, Alexandr and Bau, Claiton H.D. and Baumeister, Sarah and Baur-Streubel, Ramona and Bellgrove, Mark A. and Biederman, Joseph and Bralten, Janita and Brandeis, Daniel and Brem, Silvia and Buitelaar, Jan K. and Busatto, Geraldo F. and Castellanos, Francisco X. and Cercignani, Mara and Chaim-Avancini, Tiffany M. and Chantiluke, Kaylita C. and Christakou, Anastasia and Coghill, David and Conzelmann, Annette and Cubillo, Ana I. and Cupertino, Renata B. and de Zeeuw, Patrick and Doyle, Alysa E. and Durston, Sarah and Earl, Eric A. and Epstein, Jeffery N. and Ethofer, Thomas and Fair, Damien A. and Fallgatter, Andreas J. and Faraone, Stephen V. and Frodl, Thomas and Gabel, Matt C. and Gogberashvili, Tinatin and Grevet, Eugenio H. and Haavik, Jan and Harrison, Neil A. and Hartman, Catharina A. and Heslenfeld, Dirk J. and Hoekstra, Pieter J. and Hohmann, Sarah and H{\o}vik, Marie F. and Jernigan, Terry L. and Kardatzki, Bernd and Karkashadze, Georgii and Kelly, Clare and Kohls, Gregor and Konrad, Kerstin and Kuntsi, Jonna and Lazaro, Luisa and Lera-Miguel, Sara and Lesch, Klaus-Peter and Louza, Mario R. and Lundervold, Astri J. and Malpas, Charles B and Mattos, Paulo and McCarthy, Hazel and Namazova-Baranova, Leyla and Nicolau, Rosa and Nigg, Joel T. and Novotny, Stephanie E. and Oberwelland Weiss, Eileen and O'Gorman Tuura, Ruth L. and Oosterlaan, Jaap and Oranje, Bob and Paloyelis, Yannis and Pauli, Paul and Picon, Felipe A. and Plessen, Kerstin J. and Ramos-Quiroga, J. Antoni and Reif, Andreas and Reneman, Liesbeth and Rosa, Pedro G.P. and Rubia, Katya and Schrantee, Anouk and Schweren, Lizanne J.S. and Seitz, Jochen and Shaw, Philip and Silk, Tim J. and Skokauskas, Norbert and Soliva Vila, Juan C. and Stevens, Michael C. and Sudre, Gustavo and Tamm, Leanne and Tovar-Moll, Fernanda and van Erp, Theo G.M. and Vance, Alasdair and Vilarroya, Oscar and Vives-Gilabert, Yolanda and von Polier, Georg G. and Walitza, Susanne and Yoncheva, Yuliya N. and Zanetti, Marcus V. and Ziegler, Georg C. and Glahn, David C. and Jahanshad, Neda and Medland, Sarah E. and Thompson, Paul M. and Fisher, Simon E. and Franke, Barbara and Francks, Clyde}, title = {Analysis of structural brain asymmetries in attention-deficit/hyperactivity disorder in 39 datasets}, series = {Journal of Child Psychology and Psychiatry}, volume = {62}, journal = {Journal of Child Psychology and Psychiatry}, number = {10}, doi = {10.1111/jcpp.13396}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-239968}, pages = {1202 -- 1219}, year = {2021}, abstract = {Objective Some studies have suggested alterations of structural brain asymmetry in attention-deficit/hyperactivity disorder (ADHD), but findings have been contradictory and based on small samples. Here, we performed the largest ever analysis of brain left-right asymmetry in ADHD, using 39 datasets of the ENIGMA consortium. Methods We analyzed asymmetry of subcortical and cerebral cortical structures in up to 1,933 people with ADHD and 1,829 unaffected controls. Asymmetry Indexes (AIs) were calculated per participant for each bilaterally paired measure, and linear mixed effects modeling was applied separately in children, adolescents, adults, and the total sample, to test exhaustively for potential associations of ADHD with structural brain asymmetries. Results There was no evidence for altered caudate nucleus asymmetry in ADHD, in contrast to prior literature. In children, there was less rightward asymmetry of the total hemispheric surface area compared to controls (t = 2.1, p = .04). Lower rightward asymmetry of medial orbitofrontal cortex surface area in ADHD (t = 2.7, p = .01) was similar to a recent finding for autism spectrum disorder. There were also some differences in cortical thickness asymmetry across age groups. In adults with ADHD, globus pallidus asymmetry was altered compared to those without ADHD. However, all effects were small (Cohen's d from -0.18 to 0.18) and would not survive study-wide correction for multiple testing. Conclusion Prior studies of altered structural brain asymmetry in ADHD were likely underpowered to detect the small effects reported here. Altered structural asymmetry is unlikely to provide a useful biomarker for ADHD, but may provide neurobiological insights into the trait.}, language = {en} } @article{ScognamiglioCabezasWallscheidThieretal.2016, author = {Scognamiglio, Roberta and Cabezas-Wallscheid, Nina and Thier, Marc Christian and Altamura, Sandro and Reyes, Alejandro and Prendergast, {\´A}ine M. and Baumg{\"a}rtner, Daniel and Carnevalli, Larissa S. and Atzberger, Ann and Haas, Simon and von Paleske, Lisa and Boroviak, Thorsten and W{\"o}rsd{\"o}rfer, Philipp and Essers, Marieke A. G. and Kloz, Ulrich and Eisenman, Robert N. and Edenhofer, Frank and Bertone, Paul and Huber, Wolfgang and van der Hoeven, Franciscus and Smith, Austin and Trumpp, Andreas}, title = {Myc depletion induces a pluripotent dormant state mimicking diapause}, series = {Cell}, volume = {164}, journal = {Cell}, number = {4}, doi = {10.1016/j.cell.2015.12.033}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-190868}, pages = {668-680}, year = {2016}, abstract = {Mouse embryonic stem cells (ESCs) are maintained in a naive ground state of pluripotency in the presence of MEK and GSK3 inhibitors. Here, we show that ground-state ESCs express low Myc levels. Deletion of both c-myc and N-myc (dKO) or pharmacological inhibition of Myc activity strongly decreases transcription, splicing, and protein synthesis, leading to proliferation arrest. This process is reversible and occurs without affecting pluripotency, suggesting that Myc-depleted stem cells enter a state of dormancy similar to embryonic diapause. Indeed, c-Myc is depleted in diapaused blastocysts, and the differential expression signatures of dKO ESCs and diapaused epiblasts are remarkably similar. Following Myc inhibition, pre-implantation blastocysts enter biosynthetic dormancy but can progress through their normal developmental program after transfer into pseudo-pregnant recipients. Our study shows that Myc controls the biosynthetic machinery of stem cells without affecting their potency, thus regulating their entry and exit from the dormant state.}, language = {en} } @article{MemmelSisarioZimmermannetal.2020, author = {Memmel, Simon and Sisario, Dmitri and Zimmermann, Heiko and Sauer, Markus and Sukhorukov, Vladimir L. and Djuzenova, Cholpon S. and Flentje, Michael}, title = {FocAn: automated 3D analysis of DNA repair foci in image stacks acquired by confocal fluorescence microscopy}, series = {BMC Bioinformatics}, volume = {21}, journal = {BMC Bioinformatics}, doi = {10.1186/s12859-020-3370-8}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-229023}, year = {2020}, abstract = {Background Phosphorylated histone H2AX, also known as gamma H2AX, forms mu m-sized nuclear foci at the sites of DNA double-strand breaks (DSBs) induced by ionizing radiation and other agents. Due to their specificity and sensitivity, gamma H2AX immunoassays have become the gold standard for studying DSB induction and repair. One of these assays relies on the immunofluorescent staining of gamma H2AX followed by microscopic imaging and foci counting. During the last years, semi- and fully automated image analysis, capable of fast detection and quantification of gamma H2AX foci in large datasets of fluorescence images, are gradually replacing the traditional method of manual foci counting. A major drawback of the non-commercial software for foci counting (available so far) is that they are restricted to 2D-image data. In practice, these algorithms are useful for counting the foci located close to the midsection plane of the nucleus, while the out-of-plane foci are neglected. Results To overcome the limitations of 2D foci counting, we present a freely available ImageJ-based plugin (FocAn) for automated 3D analysis of gamma H2AX foci in z-image stacks acquired by confocal fluorescence microscopy. The image-stack processing algorithm implemented in FocAn is capable of automatic 3D recognition of individual cell nuclei and gamma H2AX foci, as well as evaluation of the total foci number per cell nucleus. The FocAn algorithm consists of two parts: nucleus identification and foci detection, each employing specific sequences of auto local thresholding in combination with watershed segmentation techniques. We validated the FocAn algorithm using fluorescence-labeled gamma H2AX in two glioblastoma cell lines, irradiated with 2 Gy and given up to 24 h post-irradiation for repair. We found that the data obtained with FocAn agreed well with those obtained with an already available software (FoCo) and manual counting. Moreover, FocAn was capable of identifying overlapping foci in 3D space, which ensured accurate foci counting even at high DSB density of up to similar to 200 DSB/nucleus. Conclusions FocAn is freely available an open-source 3D foci analyzer. The user-friendly algorithm FocAn requires little supervision and can automatically count the amount of DNA-DSBs, i.e. fluorescence-labeled gamma H2AX foci, in 3D image stacks acquired by laser-scanning microscopes without additional nuclei staining.}, language = {en} } @article{WernerWakabayashiChenetal.2019, author = {Werner, Rudolf A. and Wakabayashi, Hiroshi and Chen, Xinyu and Hayakawa, Nobuyuki and Lapa, Constantin and Rowe, Steven P. and Javadi, Mehrbod S. and Robinson, Simon and Higuchi, Takahiro}, title = {Ventricular distribution pattern of the novel sympathetic nerve PET radiotracer \(^{18}\)F-LMI1195 in Rabbit Hearts}, series = {Scientific Reports}, volume = {9}, journal = {Scientific Reports}, doi = {10.1038/s41598-019-53596-2}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-202707}, pages = {17026}, year = {2019}, abstract = {We aimed to determine a detailed regional ventricular distribution pattern of the novel cardiac nerve PET radiotracer \(^{18}\)F-LMI1195 in healthy rabbits. Ex-vivo high resolution autoradiographic imaging was conducted to identify accurate ventricular distribution of \(^{18}\)F-LMI1195. In healthy rabbits, \(^{18}\)F-LMI1195 was administered followed by the reference perfusion marker \(^{201}\)Tl for a dual-radiotracer analysis. After 20 min of \(^{18}\)F-LMI1195 distribution time, the rabbits were euthanized, the hearts were extracted, frozen, and cut into 20-μm short axis slices. Subsequently, the short axis sections were exposed to a phosphor imaging plate to determine \(^{18}\)F-LMI1195 distribution (exposure for 3 h). After complete \(^{18}\)F decay, sections were re-exposed to determine 201Tl distribution (exposure for 7 days). For quantitative analysis, segmental regions of Interest (ROIs) were divided into four left ventricular (LV) and a right ventricular (RV) segment on mid-ventricular short axis sections. Subendocardial, mid-portion, and subepicardial ROIs were placed on the LV lateral wall. \(^{18}\)F-LMI1195 distribution was almost homogeneous throughout the LV wall without any significant differences in all four LV ROIs (anterior, posterior, septal and lateral wall, 99 ± 2, 94 ± 5, 94 ± 4 and 97 ± 3\%LV, respectively, n.s.). Subepicardial \(^{201}\)Tl uptake was significantly lower compared to the subendocardial portion (subendocardial, mid-portion, and subepicardial activity: 90 ± 3, 96 ± 2 and *80 ± 5\%LV, respectively, *p < 0.01 vs. mid-portion). This was in contradistinction to the transmural wall profile of \(^{18}\)F-LMI1195 (90 ± 4, 96 ± 5 and 84 ± 4\%LV, n.s.). A slight but significant discrepant transmural radiotracer distribution pattern of \(^{201}\)Tl in comparison to \(^{18}\)F-LMI1195 may be a reflection of physiological sympathetic innervation and perfusion in rabbit hearts.}, language = {en} } @article{UhlerRedlichZhangetal.2021, author = {Uhler, Johannes and Redlich, Sarah and Zhang, Jie and Hothorn, Torsten and Tobisch, Cynthia and Ewald, J{\"o}rg and Thorn, Simon and Seibold, Sebastian and Mitesser, Oliver and Morin{\`e}re, J{\´e}r{\^o}me and Bozicevic, Vedran and Benjamin, Caryl S. and Englmeier, Jana and Fricke, Ute and Ganuza, Cristina and Haensel, Maria and Riebl, Rebekka and Rojas-Botero, Sandra and Rummler, Thomas and Uphus, Lars and Schmidt, Stefan and Steffan-Dewenter, Ingolf and M{\"u}ller, J{\"o}rg}, title = {Relationships of insect biomass and richness with land use along a climate gradient}, series = {Nature Communications}, volume = {12}, journal = {Nature Communications}, number = {1}, doi = {10.1038/s41467-021-26181-3}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-265058}, year = {2021}, abstract = {Recently reported insect declines have raised both political and social concern. Although the declines have been attributed to land use and climate change, supporting evidence suffers from low taxonomic resolution, short time series, a focus on local scales, and the collinearity of the identified drivers. In this study, we conducted a systematic assessment of insect populations in southern Germany, which showed that differences in insect biomass and richness are highly context dependent. We found the largest difference in biomass between semi-natural and urban environments (-42\%), whereas differences in total richness (-29\%) and the richness of threatened species (-56\%) were largest from semi-natural to agricultural environments. These results point to urbanization and agriculture as major drivers of decline. We also found that richness and biomass increase monotonously with increasing temperature, independent of habitat. The contrasting patterns of insect biomass and richness question the use of these indicators as mutual surrogates. Our study provides support for the implementation of more comprehensive measures aimed at habitat restoration in order to halt insect declines.}, language = {en} } @article{DjuzenovaFiedlerMemmeletal.2019, author = {Djuzenova, Cholpon S. and Fiedler, Vanessa and Memmel, Simon and Katzer, Astrid and Sisario, Dmitri and Brosch, Philippa K. and G{\"o}hrung, Alexander and Frister, Svenja and Zimmermann, Heiko and Flentje, Michael and Sukhorukov, Vladimir L.}, title = {Differential effects of the Akt inhibitor MK-2206 on migration and radiation sensitivity of glioblastoma cells}, series = {BMC Cancer}, volume = {19}, journal = {BMC Cancer}, doi = {10.1186/s12885-019-5517-4}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-200290}, pages = {299}, year = {2019}, abstract = {Background Most tumor cells show aberrantly activated Akt which leads to increased cell survival and resistance to cancer radiotherapy. Therefore, targeting Akt can be a promising strategy for radiosensitization. Here, we explore the impact of the Akt inhibitor MK-2206 alone and in combination with the dual PI3K and mTOR inhibitor PI-103 on the radiation sensitivity of glioblastoma cells. In addition, we examine migration of drug-treated cells. Methods Using single-cell tracking and wound healing migration tests, colony-forming assay, Western blotting, flow cytometry and electrorotation we examined the effects of MK-2206 and PI-103 and/or irradiation on the migration, radiation sensitivity, expression of several marker proteins, DNA damage, cell cycle progression and the plasma membrane properties in two glioblastoma (DK-MG and SNB19) cell lines, previously shown to differ markedly in their migratory behavior and response to PI3K/mTOR inhibition. Results We found that MK-2206 strongly reduces the migration of DK-MG but only moderately reduces the migration of SNB19 cells. Surprisingly, MK-2206 did not cause radiosensitization, but even increased colony-forming ability after irradiation. Moreover, MK-2206 did not enhance the radiosensitizing effect of PI-103. The results appear to contradict the strong depletion of p-Akt in MK-2206-treated cells. Possible reasons for the radioresistance of MK-2206-treated cells could be unaltered or in case of SNB19 cells even increased levels of p-mTOR and p-S6, as compared to the reduced expression of these proteins in PI-103-treated samples. We also found that MK-2206 did not enhance IR-induced DNA damage, neither did it cause cell cycle distortion, nor apoptosis nor excessive autophagy. Conclusions Our study provides proof that MK-2206 can effectively inhibit the expression of Akt in two glioblastoma cell lines. However, due to an aberrant activation of mTOR in response to Akt inhibition in PTEN mutated cells, the therapeutic window needs to be carefully defined, or a combination of Akt and mTOR inhibitors should be considered.}, language = {en} } @article{MaihoffSahlerSchogeretal.2023, author = {Maihoff, Fabienne and Sahler, Simone and Schoger, Simon and Brenzinger, Kristof and Kallnik, Katharina and Sauer, Nikki and Bofinger, Lukas and Schmitt, Thomas and Nooten, Sabine S. and Classen, Alice}, title = {Cuticular hydrocarbons of alpine bumble bees (Hymenoptera: Bombus) are species-specific, but show little evidence of elevation-related climate adaptation}, series = {Frontiers in Ecology and Evolution}, volume = {11}, journal = {Frontiers in Ecology and Evolution}, issn = {2296-701X}, doi = {10.3389/fevo.2023.1082559}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-304420}, year = {2023}, abstract = {Alpine bumble bees are the most important pollinators in temperate mountain ecosystems. Although they are used to encounter small-scale successions of very different climates in the mountains, many species respond sensitively to climatic changes, reflected in spatial range shifts and declining populations worldwide. Cuticular hydrocarbons (CHCs) mediate climate adaptation in some insects. However, whether they predict the elevational niche of bumble bees or their responses to climatic changes remains poorly understood. Here, we used three different approaches to study the role of bumble bees' CHCs in the context of climate adaptation: using a 1,300 m elevational gradient, we first investigated whether the overall composition of CHCs, and two potentially climate-associated chemical traits (proportion of saturated components, mean chain length) on the cuticle of six bumble bee species were linked to the species' elevational niches. We then analyzed intraspecific variation in CHCs of Bombus pascuorum along the elevational gradient and tested whether these traits respond to temperature. Finally, we used a field translocation experiment to test whether CHCs of Bombus lucorum workers change, when translocated from the foothill of a cool and wet mountain region to (a) higher elevations, and (b) a warm and dry region. Overall, the six species showed distinctive, species-specific CHC profiles. We found inter- and intraspecific variation in the composition of CHCs and in chemical traits along the elevational gradient, but no link to the elevational distribution of species and individuals. According to our expectations, bumble bees translocated to a warm and dry region tended to express longer CHC chains than bumble bees translocated to cool and wet foothills, which could reflect an acclimatization to regional climate. However, chain lengths did not further decrease systematically along the elevational gradient, suggesting that other factors than temperature also shape chain lengths in CHC profiles. We conclude that in alpine bumble bees, CHC profiles and traits respond at best secondarily to the climate conditions tested in this study. While the functional role of species-specific CHC profiles in bumble bees remains elusive, limited plasticity in this trait could restrict species' ability to adapt to climatic changes.}, language = {en} } @article{KernHaagsEggeretal.2023, author = {Kern, Christian S. and Haags, Anja and Egger, Larissa and Yang, Xiaosheng and Kirschner, Hans and Wolff, Susanne and Seyller, Thomas and Gottwald, Alexander and Richter, Mathias and de Giovannini, Umberto and Rubio, Angel and Ramsey, Michael G. and Bocquet, Fran{\c{c}}ois C. and Soubatch, Serguei and Tautz, F. Stefan and Puschnig, Peter and Moser, Simon}, title = {Simple extension of the plane-wave final state in photoemission: bringing understanding to the photon-energy dependence of two-dimensional materials}, series = {Physical Review Research}, volume = {5}, journal = {Physical Review Research}, number = {3}, doi = {10.1103/PhysRevResearch.5.033075}, url = {http://nbn-resolving.de/urn:nbn:de:bvb:20-opus-350330}, year = {2023}, abstract = {Angle-resolved photoemission spectroscopy (ARPES) is a method that measures orbital and band structure contrast through the momentum distribution of photoelectrons. Its simplest interpretation is obtained in the plane-wave approximation, according to which photoelectrons propagate freely to the detector. The photoelectron momentum distribution is then essentially given by the Fourier transform of the real-space orbital. While the plane-wave approximation is remarkably successful in describing the momentum distributions of aromatic compounds, it generally fails to capture kinetic-energy-dependent final-state interference and dichroism effects. Focusing our present study on quasi-freestanding monolayer graphene as the archetypical two-dimensional (2D) material, we observe an exemplary E\(_{kin}\)-dependent modulation of, and a redistribution of spectral weight within, its characteristic horseshoe signature around the \(\bar {K}\) and \(\bar {K´}\) points: both effects indeed cannot be rationalized by the plane-wave final state. Our data are, however, in remarkable agreement with ab initio time-dependent density functional simulations of a freestanding graphene layer and can be explained by a simple extension of the plane-wave final state, permitting the two dipole-allowed partial waves emitted from the C 2p\(_z\) orbitals to scatter in the potential of their immediate surroundings. Exploiting the absolute photon flux calibration of the Metrology Light Source, this scattered-wave approximation allows us to extract E\(_{kin}\)-dependent amplitudes and phases of both partial waves directly from photoemission data. The scattered-wave approximation thus represents a powerful yet intuitive refinement of the plane-wave final state in photoemission of 2D materials and beyond.}, language = {en} }