• Keine Ergebnisse gefunden

easy removal

b

AL PC TC

dorsal

anterior

outgrowth. To evaluate the locust embryo as a test system for such chemicals, the accepted test compound rotenone, inhibiting axon elongation, as well as endpoint specific controls for cytoskeletal dynamics (colchicine and cytochalasine D) and calcium signaling (verapamil and diltiazem) were used. I focused on the effects of the calcium channel blockers and the effects of altered calcium levels on Ti1 pioneers.

Intracellular calcium concentrations change considerably throughout Ti1 pioneer neuron development. Using intracellular injection of calcium indicators, Bentley et al. (1991) could show that guidepost cells are contacted by Ti1 pioneers and cause a downregulation of intracellular calcium levels. Lau et al. (1999) have demonstrated that local increase of intracellular calcium in locust growth cones promotes the generation of filopodia. Intracellular calcium can be elevated either through the release of intracellular calcium stores or by the influx through ion channels (Berridge 1997, Walz et al. 1995, Lohr 2003). Continuous blockage of ion channels that allow calcium influx, such as voltage gated calcium channels, may lead to overall decreased intracellular calcium concentrations. Verapamil and diltiazem, two classes of L-Type calcium channel blockers that are also effective in insects (Lohr et al. 2005), inhibited Ti1 outgrowth in a concentration dependent manner, with neurite outgrowth being affected at lower concentrations than general viability. Since two different pharmacological classes of channel blockers were used, unspecific actions of both blockers on pioneer outgrowth seem to be unlikely. The gathered data implies that neurite elongation may be dependent on intracellular baseline calcium concentrations, given that blockage of calcium channels does indeed reduce calcium levels in the growth cone. Since localized calcium increase in the growth cone facilitates the formation of filopodia, reducing extracellular calcium influx may in turn reduce growth speed, as fewer filopodia are formed by the growth cone during navigation.

Intracellular calcium concentrations can be increased by the application of high concentrations of caffeine. This effect is caused by ryanodine receptor activation and subsequent release of intracellular calcium from the endoplasmic reticulum (Walz et al. 1995). Following the experimental protocol by Bode et al. (2020 b), dose response curves of caffeine were generated showing a concentration dependent disruption of correct pathfinding (Fig. 3). In contrast to experiments with calcium channel blockers, caffeine had no effects on axon elongation and general viability at tested concentrations (Fig. 3 A). However, pathfinding and fasciculation of

Isbister et al. (1999) where immunological block of Sema1 resulted in a similar phenotype.

Bentley et al. (1991) have shown that intracellular calcium concentrations are decreased in the Ti1 axon in the trochanter segment, while the growth cones travel along the band of Sema1 (Kolodkin et al. 1992). As there are currently no direct calcium imaging data of Ti1 growth cones available, it is unknown whether this decrease in calcium concentration in the axon is indicative for the calcium concentration in the growth cone. However, if this decrease extends

Fig. 3 Caffeine induces pathfinding errors in Ti1 pioneer neurons of the developing locust hindleg. Caffeine is known for its depleting effect on intracellular calcium stores at high concentrations. According to the experimental protocol described by Bode et al. (2020 b), I generated dose response curves (A) of two viability (resazurine and dead-cell protease assay), and two neurite growth endpoints (elongation score and correct pathways). n indicates the number of evaluated embryos. Viability and elongation were not affected at any tested concentration. However, caffeine induced pathfinding errors were observed (IC50: 20 mM).

Ti1 pioneer neurons develop stereotypically in embryo culture (B) as described in Bergmann et al. (2019). Incubation of developing embryos with caffeine disrupted correct neurite growth in a concentration dependent manner (C & D dented arrowheads). Note that increasing caffeine concentrations (C & D) affect the quantity of erroneous growth over all specimen and not the severity. Soma marked by asterisk, end of axon marked by filled arrowhead.

Scalebar = 100 µm.

decrease of intracellular calcium leads to reduced neurite growth, whereas calcium increase disrupts correct pathway formation in the trochanter segment. Interestingly, these effects are not in opposition to each other, suggesting that intracellular calcium concentrations in Ti1 growth cones has to be tightly regulated to facilitate proper steering and outgrowth. Whether this regulation is necessary for Sema1 guided growth in the trochanter segment remains a challenging question for future research.

To resolve pioneer neuron elongation and pathfinding errors in the three dimensional limb, scanning laser optical tomography was employed for imaging and metric assessment of neurite length. Whereas glycerol clears embryos sufficiently for epifluorescence microscopy, scanning laser optical tomography requires even less opacity of the tissue to reduce light scattering.

A mixture of methyl salicylate and benzyl benzoate has better tissue infiltration properties due to its low viscosity, reducing light scattering considerably. However, suspending tissue in clearing agents with low viscosity led to insufficient spatial fixation, resulting in movement artifacts during image acquisition. Ultimately, embryos were cleared using CRISTAL (Curing Resin-Infiltrated Sample for Transparent Analysis with Light), which utilizes polymerization of clearing agents under UV irradiation (Kellner et al. 2016), enabling simultaneous clearing and immobilization of the specimen. This allowed detection of arsenite induced erroneous outgrowth of Ti1 pioneer axons using the SLOT method. Thus, SLOT can be used as a rapid and automated imaging tool in future studies of erroneous outgrowth in DNT assays.

To confirm the effects of pharmacological agents for manipulation of calcium homeostasis directly, monitoring of cytosolic calcium is required. In organisms like Drosophila, introduction of genetically encoded calcium sensors allows for targeted imaging of specific neurons (Fiala et al. 2002). In a historical context, this advance is based on decades of research into Drosophila genetics (Brand and Perrimon 1993, Adams et al. 2000, Rubin 2000). The genome of Locusta migratoria has only been sequenced a few years ago (Wang et al. 2014) and a first draft of the Schistocerca gregaria genome has been published very recently (Verlinden et al. 2020). Considering the enormous size of locust genomes (6.5 Gb and 8.8 Gb respectively), which are 2 to 3 times larger than the human genome and over 50 to 70 times larger than the genome of Drosophila, it comes to no surprise that genetic manipulation of these animals is extremely challenging. As of yet, no genetic tools allowing manipulations on the scale necessary to achieve reliable simultaneous calcium imaging of defined locust neurons

indicators (Isaacson and Hedwig 2017). This severely limits the number of cells that can be monitored simultaneously for changes in their intracellular calcium levels. For imaging of large cell populations, calcium indicators have to be introduced by passive uptake over the cell membrane. This is commonly achieved by binding an acetoxymethyl ester (AM) group to calcium indicators, allowing for free diffusion over the membrane. Once inside the intracellular lumen, the indicator is trapped by hydrolysis of the AM group, which in turn allows calcium dependent fluorescence of the indicator. However, calcium indicators that are delivered in such a manner are quickly compartmentalized or extruded from locust neurons. I found that the calcium indicator Cal-520 AM readily loads locust neurons and is largely resistant to extrusion and compartmentalization, allowing for investigation of calcium dynamics.

Calcium indicators like Cal-520 AM can be utilized in future research for neurotoxicity assays.

The developmental neurotoxicity assay presented by our lab (Bode et al. 2020 b) has the disadvantage that it cannot differentiate between inhibited neurite growth due to DNT effects and neurotoxic effects. Since disruption of neurite outgrowth and cell death of the developing neurons cause the formation of a functioning nervous system to fail, both effects need to be differentiated as they are based on different molecular mechanisms. Understanding the specific molecular mechanisms that are disrupted by a DNT compound is crucial for treatment of developmental neurotoxic effects. Long term changes in intracellular calcium concentrations in response to chemical exposure are indicative of neurotoxic effects (Tymianski et al. 1993).

Due to its intracellular retention, Cal-520 AM can be used as a fluorometric sensor for calcium levels to differentiate general neurotoxicity from developmental neurotoxicity in cultured locust neurons.

Using the next generation calcium indicator Cal-520 AM, I aimed to obtain live calcium imaging data of pioneer neurons during development. In order to monitor intracellular calcium levels in response to pharmacological manipulation I tried to replicate filet preparations of locust metathoracic limb buds (Lefcort and Bentley 1987, Bentley et al. 1991) for subsequent loading with Cal-520 AM. However, in contrast to experiments with Schistocerca gregaria, embryonic tissue of Locusta migratoria did not sufficiently adhere to coated cover slips for stable recording.

as in isolated neuron culture of insects (Bicker and Kreissl 1994, Campusano et al. 2007, Oertner et al. 1999). Since Cal-520 AM readily loads locust neurons, I tried to elucidate neurophysiological and neuroanatomical aspects of neurons in the olfactory system of Locusta migratoria. Since the number of antennal lobe neurons as well as its respective subpopulations of projection neurons and local interneurons were unknown for Locusta migratoria, the number of cells within the antennal lobe was ascertained. Accounting for glia (Gocht et al. 2009), the antennal lobe is estimated to be comprised of about 1000 neurons. Calculations based on diameter distributions allowed me to estimate that roughly 24 % of antennal lobe neurons are local neurons. These numbers correspond nicely to data from the locust species Schistocerca gregaria, where the antennal lobe is comprised of 1130 neurons with 26 % local neurons (Laurent 1996, Leitch and Laurent 1996). Ernst et al. (1977) reported that soma diameters of antennal lobe neurons in the dorsal frontal cap measure between 20 µm and 25 µm, whereas somata that lie more ventrally range between 10 µm and 12.5 µm.

However, I found a more gradual distribution of soma diameters between 10 µm and 35 µm, making identification of local interneurons and projections neurons based on diameter difficult.

Receiver operating characteristic analysis (ROC) can be used for distinguishing individuals in a population, based on a suitable criterion. ROC calculates the probabilities of false negative and false positive results at a given criterion threshold, dividing a population into two subpopulations. Using ROC analysis, soma diameter was found to be a suitable criterion for the prediction of projection/local neuron identity and a threshold diameter of about 20 µm was defined for projection/local neuron classification.

Antennal lobe neurons respond to cholinergic stimulation with an increase of cytosolic calcium and also show a wide range of responsiveness towards nicotinic or muscarinic stimulations.

Surprisingly, muscarinic calcium responses in projection neurons were found whereas Drosophila projection neurons were reported to be unresponsive to muscarinic stimulations (Rozenfeld et al. 2019). This range of nicotinic and muscarinic responsiveness in projection neurons could indicate the presence of physiologically distinct subpopulations of projection neurons (Croset et al. 2018), expressing different types or quantities of nicotinic and muscarinic receptors. Local interneurons were observed to respond differently to cholinergic stimulation than projection neurons, possibly as a consequence of the different role they fulfill in the olfactory circuit. Particularly GABAergic local neurons were more responsive to muscarinic stimulation than other local neurons. GABAergic local neurons appear to be essential for the

responsiveness of local interneurons was detected after nicotinic/muscarinic co-stimulation, suggesting short term memory-like increase of cholinergic responsiveness. In Drosophila, a similar mechanism seems to be at play (Rozenfeld et al. 2019). There, muscarinic receptors play an important role in the mitigation of short term depression of GABAergic local neurons in the olfactory circuit. In locusts, this effect on nicotinic responsiveness was found in all local neuron types, but not in all individual local neurons.

The identification of a calcium indicator that readily loads and is retained by locust neurons opens up a new avenue of research in the development of locust specific pesticides.

Calcium imaging can be used to identify compounds that specifically disrupt locust neurophysiology in vivo and in culture. Moreover, recent advancements in genome editing with CRISPR-Cas9 rekindled research regarding the locust olfactory system (Li et al. 2016).

Elucidation of the locust olfactory network utilizing calcium imaging will provide further insight in the network mechanisms and cellular properties of the locust olfactory system that can be exploited for locust pest control. In conclusion, locusts remain highly relevant as devastating pest insects, as a test system for developmental neurotoxicity and as a fruitful preparation for basic neurobiology.

J. C. (2000) The Genome Sequence of Drosophila melanogaster. Science 287 2185-2195.

Amar M., Thomas P., Wonnacott S., and Lunt G. G. (1995) A nicotinic acetylcholine receptor subunit from insect brain forms a non-desensitising homo-oligomeric nicotinic acetylcholine receptor when expressed in Xenopus oocytes. Neuroscience Letters 199 107-110.

Anton S. and Hansson B. S. (1996) Antennal lobe interneurons in the desert locust Schistocerca gregaria (Forskal): Processing of aggregation pheromones in adult males and females.

Journal of Comparative Neurology 370 85-96.

Bate C. M. (1976) Pioneer neurones in an insect embryo. Nature 260 54-56.

Bentley D. and Caudy M. (1983) Pioneer axons lose directed growth after selective killing of guidepost cells. Nature 304 62-65.

Bentley D., Guthrie P. B., and Kater S. B. (1991) Calcium ion distribution in nascent pioneer axons and coupled preaxonogenesis neurons in situ. The Journal of Neuroscience 11 1300-1308.

Berridge M. J. (1997) Elementary and global aspects of calcium signalling. The Journal of Physiology 499 291-306.

Bicker G. and Kreissl S. (1994) Calcium imaging reveals nicotinic acetylcholine receptors on cultured mushroom body neurons. Journal of Neurophysiology 71 808-810.

Bode K., Bohn M., Reitmeier J., Betker P., Stern M., and Bicker G. (2020 b) A locust embryo as predictive developmental neurotoxicity testing system for pioneer axon pathway formation. Archives of Toxicology 94 4099-4113.

Brand A. H. and Perrimon N. (1993) Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118 401-415.

Breer H. and Knipper M. (1984) Characterization of acetylcholine release from insect synaptosomes. Insect Biochemistry 14 337-344.

Breer H. and Sattelle D. B. (1987) Molecular properties and functions of insect acetylcholine receptors. Journal of Insect Physiology 33 771-790.

Buck L. and Axel R. (1991) A novel multigene family may encode odorant receptors: A molecular basis for odor recognition. Cell 65 175-187.

Campusano J. M., Su H., Jiang S. A., Sicaeros B., and O'Dowd D. K. (2007) nAChR-mediated calcium responses and plasticity in Drosophila Kenyon cells. Developmental Neurobiology 67 1520-1532.

Caulfield M. P. (1993) Muscarinic Receptors—Characterization, coupling and function.

Pharmacology & Therapeutics 58 319-379.

Condic M. L. and Bentley D. (1989) Pioneer neuron pathfinding from normal and ectopic locations in vivo after removal of the basal lamina. Neuron 3 427-439.

Croset V., Treiber C. D., and Waddell S. (2018) Cellular diversity in the Drosophila midbrain revealed by single-cell transcriptomics. eLife 7 doi: 10.7554/eLife.34550

Ehrhardt E. and Boyan G. (2020) Evidence for the cholinergic markers ChAT and vAChT in sensory cells of the developing antennal nervous system of the desert locust Schistocerca gregaria. Invertebrate Neuroscience 20 19 doi: 10.1007/s10158-020-00252-4

Ernst K. D., Boeckh J., and Boeckh V. (1977) A neuroanatomical study on the organization of the central antennal pathways in insects. Cell and Tissue Research 176 285-308.

FAO, Food and Agriculture Organization of the United Nations (2009) Frequently Asked Questions (FAQs) about locusts. Accessed 05 March 2021.

http://www.fao.org/ag/locusts/en/info/info/faq/index.html

Fiala A., Spall T., Diegelmann S., Eisermann B., Sachse S., Devaud J., . . . Galizia C. G. (2002) Genetically Expressed Cameleon in Drosophila melanogaster Is Used to Visualize Olfactory Information in Projection Neurons. Current Biology 12 1877-1884.

Galizia C. G. (2014) Olfactory coding in the insect brain: data and conjectures. European Journal of Neuroscience 39 1784-1795.

Gocht D., Wagner S., and Heinrich R. (2009) Recognition, presence, and survival of locust central nervous glia in situ and in vitro. Microscopy Research and Technique 72 385-397.

Gomez T. M. and Zheng J. Q. (2006) The molecular basis for calcium-dependent axon pathfinding. Nature Reviews Neuroscience 7 115-125.

Goodman C. S. and Bate M. (1981) Neuronal development in the grasshopper. Trends in Neurosciences 4 163-169.

Gundelfinger E. D. (1992) How complex is the nicotinic receptor system of insects? Trends in Neurosciences 15 206-211.

Harrelson A. L. and Goodman C. S. (1988) Growth cone guidance in insects: fasciclin II is a member of the immunoglobulin superfamily. Science 242 700-708.

Hermsen B., Stetzer E., Thees R., Heiermann R., Schrattenholz A., Ebbinghaus U., . . . Maelicke A. (1998) Neuronal nicotinic receptors in the Locust Locusta migratoria: cloning and expression. Journal of Biological Chemistry 273 18394-18404.

Isaacson M. D. and Hedwig B. (2017) Electrophoresis of polar fluorescent tracers through the nerve sheath labels neuronal populations for anatomical and functional imaging.

Scientific Reports 7 40433 doi: 10.1038/srep40433

Isbister C. M. and O'Connor T. P. (2000) Mechanisms of growth cone guidance and motility in the developing grasshopper embryo. Journal of Neurobiology 44 271-280.

Isbister C. M., Tsai A., Wong S. T., Kolodkin A. L., and O'Connor T. P. (1999) Discrete roles for secreted and transmembrane semaphorins in neuronal growth cone guidance in vivo.

Development 126 2007-2019.

Jones A. K., Marshall J., Blake A. D., Buckingham S. D., Darlison M. G., and Sattelle D. B.

(2005) Sgbeta1, a novel locust (Schistocerca gregaria) non-alpha nicotinic acetylcholine receptor-like subunit with homology to the Drosophila melanogaster Dbeta1 subunit.

Invertebrate Neuroscience 5 147-155.

Knipper M. and Breer H. (1988) Subtypes of muscarinic receptors in insect nervous system.

Comparative Biochemistry and Physiology Part C: Comparative Pharmacology 90 275-280.

Kolodkin A. L., Matthes D. J., O'Connor T. P., Patel N. H., Admon A., Bentley D., and Goodman C. S. (1992) Fasciclin IV: Sequence, expression, and function during growth cone guidance in the grasshopper embryo. Neuron 9 831-845.

Krug A. K., Balmer N. V., Matt F., Schönenberger F., Merhof D., and Leist M. (2013) Evaluation of a human neurite growth assay as specific screen for developmental neurotoxicants. Archives of Toxicology 87 2215-2231.

Kutsch W. and Bentley D. (1987) Programmed death of peripheral pioneer neurons in the grasshopper embryo. Developmental Biology 123 517-525.

Lau P., Zucker R. S., and Bentley D. (1999) Induction of Filopodia by Direct Local Elevation of Intracellular Calcium Ion Concentration. Journal of Cell Biology 145 1265-1276.

Laurent G. (1996) Dynamical representation of odors by oscillating and evolving neural assemblies. Trends in Neurosciences 19 489-496.

Lefcort F. and Bentley D. (1987) Pathfinding by pioneer neurons in isolated, opened and mesoderm-free limb buds of embryonic grasshoppers. Developmental Biology 119 466-480.

Leitch B. and Laurent G. (1996) GABAergic synapses in the antennal lobe and mushroom body of the locust olfactory system. Journal of Comparative Neurology 372 487-514.

Li Y., Zhang J., Chen D., Yang P., Jiang F., Wang X., and Kang L. (2016) CRISPR/Cas9 in locusts: Successful establishment of an olfactory deficiency line by targeting the mutagenesis of an odorant receptor co-receptor (Orco). Insect Biochemistry and Molecular Biology 79 27-35.

Lohr C. (2003) Monitoring neuronal calcium signalling using a new method for ratiometric confocal calcium imaging. Cell Calcium 34 295-303.

Lohr C., Heil J. E., and Deitmer J. W. (2005) Blockage of voltage-gated calcium signaling impairs migration of glial cells in vivo. Glia 50 198-211.

Lowery L. A. and Van Vactor D. (2009) The trip of the tip: understanding the growth cone machinery. Nature reviews Molecular cell biology 10 332-343.

MacLeod K. and Laurent G. (1996) Distinct Mechanisms for Synchronization and Temporal Patterning of Odor-Encoding Neural Assemblies. Science 274 976-979.

Marshall J., Buckingham S. D., Shingai R., Lunt G. G., Goosey M. W., Darlison M. G., . . . Barnard E. A. (1990) Sequence and functional expression of a single alpha subunit of an insect nicotinic acetylcholine receptor. The EMBO Journal 9 4391-4398.

Moreaux L. and Laurent G. (2007) Estimating firing rates from calcium signals in locust projection neurons in vivo. Frontiers in Neural Circuits 1 doi:

10.3389/neuro.04.002.2007

Nevo D. (1996) The Desert Locust, Schistocerca gregaria, and Its Control in the Land of Israel and the Near East in Antiquity, with Some reflections on Its Appearance in Israel in Modern Times. Phytoparasitica 24 7-32.

OECD, Organisation for Economic Co-operation and Development. (2006) Test No. 426:

Developmental Neurotoxicity Study. OECD Publishing, Accessed 05 March 2021.

https://www.oecd-ilibrary.org/environment/test-no-426-developmental-neurotoxicity-study_9789264067394-en

Oertner T. G., Single S., and Borst A. (1999) Separation of voltage- and ligand-gated calcium influx in locust neurons by optical imaging. Neuroscience Letters 274 95-98.

Pamies D., Block K., Lau P., Gribaldo L., Pardo C. A., Barreras P., . . . Hogberg H. T. (2018) Rotenone exerts developmental neurotoxicity in a human brain spheroid model.

Toxicology and Applied Pharmacology 354 101-114.

Pasterkamp R. J. and Kolodkin A. L. (2003) Semaphorin junction: making tracks toward neural connectivity. Current Opinion in Neurobiology 13 79-89.

Polleux F., Morrow T., and Ghosh A. (2000) Semaphorin 3A is a chemoattractant for cortical apical dendrites. Nature 404 567-573.

Rajagopalan A. and Assisi C. (2020) Effect of Circuit Structure on Odor Representation in the Insect Olfactory System. eNeuro 7 ENEURO.0130-19.2020 doi: 10.1523/eneuro.0130-19.2020

Rind F. C. and Leitinger G. (2000) Immunocytochemical evidence that collision sensing neurons in the locust visual system contain acetylcholine. Journal of Comparative Neurology 423 389-401.

Rössler W., Randolph P. W., Tolbert L. P., and Hildebrand J. G. (1999) Axons of olfactory receptor cells of transsexually grafted antennae induce development of sexually dimorphic glomeruli in Manduca sexta. Journal of Neurobiology 38 521-541.

Rozenfeld E., Lerner H., and Parnas M. (2019) Muscarinic Modulation of Antennal Lobe GABAergic Local Neurons Shapes Odor Coding and Behavior. Cell Reports 29 3253-3265.e4 doi: 10.1016/j.celrep.2019.10.125

Rubin G. M. (2000) Biological annotation of the Drosophila genome sequence. Novartis Foundation Symposium 229 79-82; discussion 82-83.

Russell W. M. S., Burch R. L., and Hume C. W. (1959). The principles of humane experimental technique, vol 238. Methuen, London.

Seidel C. and Bicker G. (1997) Colocalization of NADPH-diaphorase and GABA-immunoreactivity in the olfactory and visual system of the locust. Brain Research 769 273-280.

Smirnova L., Hogberg H. T., Leist M., and Hartung T. (2014) Developmental neurotoxicity - challenges in the 21st century and in vitro opportunities. ALTEX 31 129-156.

714.

Trimmer B. A. (1994) Characterization of a muscarinic current that regulates excitability of an identified insect motoneuron. Journal of Neurophysiology 72 1862-1873.

Trimmer B. A. (1994) Characterization of a muscarinic current that regulates excitability of an identified insect motoneuron. Journal of Neurophysiology 72 1862-1873.