MECHANISMS OF GNAO1-ASSOCIATED NEUROLOGICAL DISORDERS By Huijie Feng A DISSERTATION Michigan State University in partial fulfillment of the requirements Submitted to for the degree of Pharmacology and Toxicology—Doctor of Philosophy 2019 ABSTRACT MECHANISMS OF GNAO1-ASSOCIATED NEUROLOGICAL DISORDERS By Huijie Feng Tremendous advances in the genetics of neurodevelopmental disorders have markedly improved the understanding of disease mechanisms. This project will focus on understanding the mechanisms of GNAO1 encephalopathies, a devastating but complex disorder, which exhibit multiple neurological symptoms. These include developmental delay and variable components of early onset epilepsy and/or hyperkinetic movement disorders (MDs). These symptoms are associated with mutations in the GNAO1 gene, which encodes the Gαo protein. GNAO1 mutation-associated neurological disorders include neurodevelopmental delay with involuntary movements (NEDIM, OMIM#617493) and early infantile epileptic encephalopathy (EIEE17, OMIM#615473). The number of identified patients and mutant alleles for EIEE17 or NEDIM is increasing rapidly. Gαo is the most abundant membrane protein in the mammalian central nervous system. It couples to multiple G protein-coupled receptors (GPCRs) including GABAB, α2 adrenergic, D2 dopamine, and adenosine A1 receptors; all are associated with both MDs and epilepsy. In addition, GPCRs are readily targetable by agonists and antagonists. This provides the possibility of treating GNAO1-associated neurological disorders. This project revealed a fundamental mechanistic distinction among these GNAO1 mutations: Loss-of-function (LOF) GNAO1 alleles are associated with epilepsy, while gain-of-function (GOF) GNAO1 alleles are associated primarily with MDs. However, this simple model is insufficient to explain all clinical observations. To explore this correlation, we have created mouse models carrying two of the most common human GNAO1 mutant alleles (G203R and R209H). They largely share the human pathophysiology; the G203R mouse model exhibits both MD and enhanced seizure propensity, while the R209H mutant results in MD alone, as seen in children with those mutations. Using these models, we can further explore mechanisms that lead to distinct patterns in human GNAO1 encephalopathies. To explore electrophysiological alterations in the Gnao1 G203R mutant mouse model, I performed patch clamp studies on cerebellar Purkinje cells. The results show a decreased frequency of both miniature inhibitory postsynaptic currents (mIPSCs) and spontaneous inhibitory postsynaptic currents (sIPSCs), suggesting a reduced presynaptic GABA release. This study provides a molecular and physiological understanding of different GNAO1 alleles in vitro, and identifies key candidate alleles for further analysis in in vivo mouse models and in human GNAO1-associated neurological disorders. Furthermore, our study may serve as a prototype for other correlations between reported monogenic mutations and human neurological disorders. ACKNOWLEDGEMENTS I would like to thank my mentor Dr. Richard R. Neubig. I consider myself fortunate to have had this opportunity to work with him. Through Dr. Neubig’s mentorship I have become not just a better scientist but also a better writer and presenter. I appreciate Rick’s constant attention to details and his enthusiasm in pushing my project forward. His suggestions have been invaluable to the results presented in this dissertation. I also would like to thank him for his encouragement and trust for me to freely explore and decide what I deem the best in science and in my career development. This helps me to become an independent thinker and researcher. I also highly value the mentorship of Dr. Yukun Yuan, who has unique and extensive experience and knowledge in doing patch clamp recordings. I deeply appreciate his generous sharing of his research equipment and his help in troubleshooting failed experiments. My other committee members have also played an important role in guiding the development of the work presented in this dissertation. I would like to thank Dr. Colleen Hegg, Dr. James Galligan and Dr. Charles Lee Cox for their insights and suggestions during every committee meeting I had. Their questions and perspectives have helped me assess my experimental results from alternative viewpoints and have enhanced the work presented here. I would also like to thank any clinical knowledge and detailed explanations regarding movement disorders and epilepsy given by Dr. Christos Sidiropoulos and Dr. Florian Kagerer. iv ! Within the Neubig lab, I would like to thank Jeff Leipprandt who managed with all my orders and animal subjects. In addition, I would also like to thank former research associate Dr. Benita Sjögren, who patiently helped me through my first year in the lab and whose detailed lab records provided me invaluable resources when in need. I am thankful for our Masters student Cassie Larrivee, who helped with a majority of behavioral experiments. I also would like to express my gratitude to the current members of the laboratory. Dr. Erika Lisabeth, Behirda Karaj, Dr. Vincent Shaw, Sean Misek, Yajing Ji, and Zhe Zhang have been great friends and colleagues during my time in the lab. I also value the assistance of our undergraduate student, Nils Wellhausen, who has been important to my ability to conduct this research. I would like to thank the MSU Transgenic Core headed by Dr. Elena Demireva and Dr. Huirong Xie. Their work provided me with animal subjects to work with. I also would like to thank Dr. Thomas Dexheimer from the Assay Development and Drug Repurposing Core, who gave me valuable suggestions for my experiments. I would like to thank the Department of Pharmacology and Toxicology. Dr. Anne Dorrance offered mentorship and support during my time as a graduate student at MSU. In addition, I would like to thank the department office staff for their patience and assistance with administrative issues during my completion of the department’s academic requirements. The research presented in this dissertation would not have been possible without the v !! funding received from numerous sources. I would like to express my thanks to the Bow Foundation, whose grants supported my research, and to American Epilepsy Society (AES), for the pre-doctoral fellowship that supplied my stipend for a year. I would also like to thank the Neuroscience Program for providing me with a teaching assistantship and the Department of Pharmacology & Toxicology for a dissertation completion fellowship. I have also received financial support from the AES, Society of Neurosciences (SfN) and MSU Council of Graduate Students to attend scientific meetings. A heartfelt thanks also goes to all the GNAO1 patients and families. Their support is a big driving force of the advancement of this project. I sincerely hope this project will benefit the development of new therapeutics for treating the GNAO1-encephalopathies. I reserve special thanks for my friends and family. I owe them a great deal of credit for my success. vi ! TABLE OF CONTENTS LIST OF TABLES ............................................................................................................ xi LIST OF FIGURES ........................................................................................................ xiii KEY TO ABBREVIATIONS .......................................................................................... xviii CHAPTER 1: AN OVERVIEW OF GNAO1-ASSOCIATED NEUROLOGICAL DISORDERS .................................................................................................................... 1 1.1 Abstract ................................................................................................................... 2 1.2 Introduction ............................................................................................................. 2 1.3 Gαo (GNAO1) mechanisms ..................................................................................... 5 1.4 The clinical spectrum of GNAO1 mutation-associated movement disorders .......... 7 1.4.1 GNAO1 encephalopathy displays a variety of neurological symptoms ............ 7 1.4.2 Movement disorders related to GNAO1 encephalopathy show limited response to pharmacological treatments ................................................................... 9 1.5 Potential pathogenic mechanisms of GNAO1-associated movement disorders ... 13 1.5.1 Role of cAMP regulation in movement disorders ........................................... 14 1.5.1.1 GNAL ....................................................................................................... 17 1.5.1.2 GNB1 ....................................................................................................... 18 1.5.1.3 ADCY5 ..................................................................................................... 19 1.5.1.4 PDE10A ................................................................................................... 21 1.5.1.5 Summary of cAMP regulation in movement control ................................. 22 1.5.2 Role of neurotransmitter release and synaptic vesicle fusion in movement disorders .................................................................................................................. 23 1.5.2.1 SYT1 ........................................................................................................ 25 1.5.2.2 PRRT2 ..................................................................................................... 26 1.5.2.3 SNAP25 ................................................................................................... 27 1.5.2.4 KCNMA1 .................................................................................................. 28 1.5.2.5 CACNA1A ................................................................................................ 29 1.5.2.6 CACNA1B ................................................................................................ 30 1.5.2.7 GPR88 ..................................................................................................... 31 1.5.2.8 Summary of neurotransmitter release and synaptic vesicle fusion in movement control ................................................................................................ 32 1.6 Developmental defects may also contribute to movement disorders .................... 32 1.7 Conclusion ............................................................................................................ 34 1.8 Organization of the thesis ..................................................................................... 35 APPENDIX .................................................................................................................. 37 REFERENCES ............................................................................................................ 46 vii ! CHAPTER 2: MOVEMENT DISORDER IN GNAO1 ENCEPHALOPATHY ASSOCIATED WITH GAIN-OF-FUNCTION MUTATIONS ............................................ 66 2.1 Abstract ................................................................................................................. 67 2.2 Introduction ........................................................................................................... 68 2.3 Materials and Methods .......................................................................................... 70 2.3.1 Materials ......................................................................................................... 70 2.3.2 DNA constructs and mutagenesis ................................................................ 70 2.3.3 Cell culture and transfections ....................................................................... 71 2.3.4 SDS-PAGE and Western blot ....................................................................... 72 2.3.5 cAMP measurements .................................................................................. 73 2.3.6 Data analysis and statistics ......................................................................... 73 2.4 Results .................................................................................................................. 74 2.4.1 Most pathogenic GNAO1 mutations cause reduced Gαo protein expression ................................................................................................................................. 74 2.4.2 Validation of an in vitro assay to assess function of GNAO1 mutations ......... 76 2.4.3 Nine GNAO1 mutations result in loss- (LOF) or partial loss-of-function (PLOF) ................................................................................................................................. 77 2.4.4 Six GNAO1 mutations result in gain-of-function (GOF) or normal function (NF) ................................................................................................................................. 80 2.4.5 Clinical correlation with biochemical behavior of mutant GNAO1 alleles ....... 86 2.4.6 Location of mutations linked to GNAO1 encephalopathies in the Gαo protein ................................................................................................................................. 87 2.5 Discussion ............................................................................................................. 88 APPENDICES ............................................................................................................. 93 APPENDIX A: SUPPLEMENTAL DATA .................................................................. 94 APPENDIX B: VALIDATION OF THE GENOTYPE-PHENOTYPE CORRELATION OF GNAO1-ASSOCIATED NEUROLOGICAL DISORDERS ................................ 103 REFERENCES .......................................................................................................... 112 CHAPTER 3: BEHAVIORAL ASSESSMENT OF MOUSE MODELS WITH GNAO1-ASSOCIATED MOVEMENT DISORDER AND EPILEPSY ........................... 119 3.1 Abstract ............................................................................................................... 120 3.2 Introduction ......................................................................................................... 121 3.3 Materials and Methods ........................................................................................ 126 3.3.1 Animals ......................................................................................................... 126 3.3.2 Generation of Gnao1 mutant mice ............................................................... 126 3.3.2.1 Generation of Gnao1+/G203R mouse model ............................................. 126 3.3.2.2 Generation of Gnao1+/R209H mouse model ............................................. 130 3.3.3 Genotyping and Breeding ............................................................................. 132 3.3.3.1 Genotyping Gnao1+/G203R mice .............................................................. 132 3.3.3.2 Genotyping Gnao1+/R209H mice ............................................................... 133 3.3.4 Behavioral Studies ........................................................................................ 134 viii ! 3.3.4.1 Open Field ............................................................................................. 134 3.3.4.2 RotaRod ................................................................................................. 135 3.3.4.3 Grip Strength .......................................................................................... 135 3.3.4.4 DigiGait .................................................................................................. 136 3.3.5 PTZ Kindling Susceptibility ........................................................................... 137 3.3.6 Data Analysis ................................................................................................ 137 3.4 Results ................................................................................................................ 138 3.4.1 The growth patterns of the three newly developed Gnao1 mutant mouse models (G203R, R209H and ΔT191F197) ............................................................ 138 3.4.1.1 Gnao1+/G203R mice showed normal viability and growth. ........................ 138 3.4.1.2 Gnao1+/R209H mice have expected frequency and normal viability ......... 138 3.4.1.3 Gnao1+/ΔT191F197 mice developed spontaneous seizures at P7 and died before P16 ......................................................................................................... 139 3.4.2 Behavioral assessment of mutant Gnao1 mouse models (G184S, G203R, R209H and KO) for movement patterns ................................................................ 141 3.4.2.1 Female Gnao1+/G184S and male Gnao1+/G203R mice show similar movement abnormalities and gait disturbances .................................................................. 141 3.4.2.2 Gnao1+/R209H mouse model exhibits unique hyperactive behavior in the open field arena but no abnormal motor coordination in other behavior tests and minor disturbance in gait analysis ...................................................................... 149 3.4.2.3 Previously described Gnao1+/- mouse model did not show any abnormalities in the behavioral test battery ....................................................... 152 3.4.3 PTZ kindling study of G203R and R209H mice. ........................................... 155 3.4.3.1 Male Gnao1+/G203R mice are sensitized to PTZ kindling. ........................ 155 3.4.3.2 R209H mutant mice are not hypersensitive to PTZ kindling .................. 157 3.5 Discussion ........................................................................................................... 158 APPENDIX ................................................................................................................ 165 REFERENCES .......................................................................................................... 183 CHAPTER 4: MICE WITH GNAO1-ASSOCIATED MOVEMENT DISORDER EXHIBIT REDUCED INHIBITORY SYNAPTIC INPUT TO CEREBELLAR PURKINJE CELLS ...... ...................................................................................................................................... 193 4.1 Abstract ............................................................................................................... 194 4.2 Introduction ......................................................................................................... 195 4.3 Materials and Methods. ....................................................................................... 198 4.3.1 Tissue preparation and solutions .............................................................. 198 4.3.2 Electrophysiological recording .................................................................. 199 4.3.3 Pharmacology ........................................................................................... 201 4.3.4 SDS Page and Western Blots ................................................................... 201 4.3.5 Statistical Analysis .................................................................................... 203 4.4 Results ................................................................................................................ 204 4.4.1 Presynaptic GABA release is suppressed in the cerebellar Purkinje cells of ix ! Gnao1+/G203R mice .............................................................................................. 205 4.4.2 Gαo blockers can reverse the enhanced inhibition of mIPSC frequency in G203R mice ....................................................................................................... 207 4.4.3 Presynaptic glutamate release is not affected by the G203R mutation in Gαo protein ................................................................................................................ 211 4.4.4 Effects of G-protein coupled GABAB receptors on AP-independent GABA release onto Purkinje cells ................................................................................. 213 4.4.5 Effects of G-protein coupled α2A adrenergic receptors on AP-dependent GABA release onto Purkinje cells ...................................................................... 215 4.4.6 α2A adrenergic receptor-induced inhibition of sIPSC frequency depends on activation of voltage-gated calcium channels .................................................... 217 4.4.7 G203R mice exhibit decreased Gαo protein expression but no change in Gβ levels .................................................................................................................. 221 4.5 Discussion ........................................................................................................... 223 APPENDIX ................................................................................................................ 230 REFERENCES .......................................................................................................... 236 CHAPTER 5: CONCLUSION AND FUTURE DIRECTIONS ....................................... 244 5.1 General conclusion ............................................................................................. 243 5.2 Testing the functional changes of a growing variety of GNAO1 mutations ......... 248 5.2.1 Do GNAO1 mutations affect Go’s inhibition of high-voltage gated calcium channels (N- type & P/Q- type calcium channels)? ............................................... 248 5.2.2 Do GNAO1 mutations affect Go’s activation of G protein-regulated inward rectifying potassium (GIRK) channels? ................................................................. 251 5.2.3 How do GNAO1 mutations affect G protein-regulated neurite outgrowth? .. 252 5.3 Comparison between R209H and G203R mouse models .................................. 254 5.3.1 Do G203R and R209H mouse models exhibit delayed development? ........ 255 5.3.2 How does the G203R mutation in Gαo lead to epileptogenesis? ................. 258 5.3.3 How do G203R and R209H mice differ in movement disorder phenotypes? ............................................................................................................................... 260 5.3.4 How is sex involved in abnormalities of the Gnao1 mutant mice? ............... 262 5.4 Development of a high-throughput assay for drug repurposing or drug development .............................................................................................................. 267 5.5 Impact of work in this thesis on the field ............................................................. 270 APPENDIX ................................................................................................................ 272 REFERENCES .......................................................................................................... 281 x !! LIST OF TABLES Table 1.1 Most Common GNAO1 Mutant Alleles Associated With Movement Disorders 8 Table 1.2 Drugs showing beneficial effects to control involuntary movements of GNAO1 encephalopathy patients ................................................................................................... 9 Table 1.3 Patients responding positively to deep brain stimulation (DBS) ..................... 12 Table S1.1 A complete summary of clinical information regarding GNAO1 patients ..... 38 Table 2.1 Functional data for loss-of-function (LOF) and partial loss-of-function (PLOF) mutants. .......................................................................................................................... 79 Table 2.2 Functional data for normal and gain-of-function (GOF) mutants. ................... 82 Table 2.3 Correlation between cAMP inhibition and clinical diagnosis. .......................... 83 Table S2.1 Primer sequences for mutagenesis to create GNAO1 mutants. .................. 95 Table S2.2 All non-functioning GNAO1 mutations tested with GNAS KO cells ............ 110 Table S2.3 All functioning GNAO1 mutations tested with GNAS KO cells ................... 110 Table S2.4 Genotype-phenotype correlation of the newly reported GNAO1 mutations ...................................................................................................................................... 111 Table 3.1 The status of Gnao1 mutant mice ................................................................. 125 Table 3.2 Location, sequence and genotyping of gRNA targets in the Gnao1 locus. .. 128 Table 3.3 Location and sequence of gRNA and ssODN template for CRISPR-Cas targeting Gnao1 locus; primers and genotyping method for Gnao1+/R209H mice ........... 131 Table 3.4 Phenotypes of Gnao1 mutant mice .............................................................. 160 xi ! Table S3.1 Gait analysis parameters of male Gnao1 G203R mutant mice ................ 173 Table S3.2 Gait analysis parameters of female Gnao1 G203R mutant mice ............. 174 Table S3.3 Gait analysis parameters of male Gnao1 G184S mutant mice ................ 175 Table S3.4 Gait analysis parameters of female Gnao1 G184S mutant mice ............. 176 Table S3.5 Gait analysis parameters of male Gnao1 R209H mutant mice .................. 177 Table S3.6 Gait analysis parameters of female Gnao1 R209H mutant mice ............. 178 Table S3.7 Gait analysis parameters of male Gnao1 KO mutant mice ........................ 179 Table S3.8 Gait analysis parameters of female Gnao1 KO mutant mice ................... 180 Table S3.9 Benchling off-target list for Gnao1 G203R gRNA ..................................... 181 Table 5.1 Comparison of clinical patterns of G203R and R209H patients ................. 254 xii ! LIST OF FIGURES Figure 1.1 Summary of reported cases of GNAO1 encephalopathy ................................ 5 Figure 1.2 Genes regulating the cAMP pathway are related to movement disorders .... 16 Figure 1.3 Pathogenic mutations in genes that regulate neurotransmitter release ........ 25 Figure S1.1 Correlation between seizure frequency and a severe EEG/MRI result ....... 44 Figure 2.1 Location and protein expression levels of human GNAO1 mutations related to epileptic encephalopathy. ............................................................................................... 75 Figure 2.2 Effect on α2A AR-mediated cAMP inhibition by GNAO1 mutants. ................ 78 Figure S2.1 Alignment of the human and mouse Gαo protein sequences. ..................... 97 Figure S2.2 Validation of the Lance Ultra cAMP assay with transient transfection of α2A adrenergic receptor (α2A AR) and Pertussis toxin (PTX)-insensitive Gαo ....................... 98 Figure S2.3 Assessment of dominant-negative effect of complete LOF mutants G40R, N270H, D174G, T191_F197del, L199P, and F275S. ..................................................... 99 Figure S2.4 Paired t-test analysis of LogEC50 values for normal and GOF mutants .... 101 Figure S2.5 Mapping of mutations on the structure of Gαo-GDP bound to RGS16 ...... 102 Figure S2.6 Comparison of validation of the Lance Ultra cAMP assay between HEK293T and GNAS KO HEK293 cells with transient transfection of α2A adrenergic receptor (α2AR) and Pertussis toxin (PTX)-insensitive Gαo .................................................................... 107 Figure S2.7 GNAO1 mutations’ functionalities correlate to their protein expression patterns. ........................................................................................................................ 108 Figure S2.8 GNAO1 mutations’ functionalities correlate to their protein expression patterns (assays done by Nils Wellhausen with a different group of mutations from those xiii ! in Figure S2.7). ............................................................................................................. 109 Figure 3.1 Development of Gnao1+/G203R mouse model. .............................................. 127 Figure 3.2 Targeting of the mouse Gnao1 locus. ......................................................... 131 Figure 3.3 The timeline for utilizing animals in this study. ............................................ 134 Figure 3.4 Gnao1+/ΔT191F197 mice developed spontaneous seizures at P7 and died before P16. .............................................................................................................................. 140 Figure 3.5 Female Gnao1+/G184S mice and male Gnao1+/G203R mice show reduced time on RotaRod and reduced grip strength. ............................................................................. 143 Figure 3.6 G184S mutant mice showed reduced activities in Open Field Test but G203R mutants do not. ............................................................................................................. 145 Figure 3.7 DigiGait Imaging System reveals sex-specific gait abnormalities in Gnao1+/G184S mice and Gnao1+/G203R mice. ................................................................... 148 Figure 3.8 Gnao1+/R209H mice show significant hyperactivity and reduced time in center in the open field arena. ................................................................................................... 150 Figure 3.9. Male and female Gnao1+/R209H mice shows gait abnormalities in different tests on the DigiGait imaging system. .......................................................................... 151 Figure 3.10 Male and female Gnao1+/- mice do not show any abnormalities in the behavioral tests including open field, Rotarod, and grip strength. ................................ 154 Figure 3.11 DigiGait Imaging System reveals the decreased stride length in male Gnao1+/- mice. .............................................................................................................. 155 Figure 3.12 Gnao1+/G203R male mice have an enhanced Pentylenetetrazol (PTZ) Kindling response. .................................................................................................................... 157 Figure 3.13 Gnao1+/R209H mice do not have an enhanced pentylenetetrazol (PTZ) kindling response .......................................................................................................... 158 xiv ! Figure S3.1 RotaRod test was conducted with 5 training sessions and 1 test session over two consecutive days. ........................................................................................... 166 Figure S3.2 Time spent at the center in the Open Field Test. ...................................... 167 Figure S3.3 RotaRod learning curve was collected in 10 consecutive tests with a 5-min break between each test. ............................................................................................ 168 Figure S3.4 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/G184S mice. ............................................. 169 Figure S3.5 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/G203R mice. ............................................. 170 Figure S3.6 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/R209H mice. ........................................... 171 Figure S3.7 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/- mice. ................................................... 172 Figure 4.1 Cerebellar Purkinje cells in brain slices from 4-6 week-old G203R mice display reduced GABAergic spontaneous synaptic currents (sIPSCs) and reduced miniature synaptic currents (mIPSCs). ......................................................................... 206 Figure 4.2 The frequencies of mIPSCs were sensitive to NEM, an inhibitor of Gαi/o proteins. ........................................................................................................................ 209 Figure 4.3 A selective inhibitor of Gi/o, pertussis toxin (PTX), increased the frequency of mIPSCs in G203R but not WT mice. ............................................................................ 210 Figure 4.4 G203R mutant slices show no difference in either spontaneous excitatory postsynaptic currents (sEPSCs) or miniature excitatory postsynaptic currents (mIPSCs). ...................................................................................................................................... 212 Figure 4.5 Activating GABAB receptor with baclofen reduces mIPSC frequency but not amplitude. PTX incubation eliminates baclofen-induced inhibition of mIPSC frequency in WT and G203R mice. ................................................................................................... 214 Figure 4.6 The frequency of sIPSCs is modulated by α2AR receptors. ......................... 216 xv ! Figure 4.7 Cadmium-block of extracellular calcium influx suppresses the frequency of sIPSCs in both WT and G203R mice. .......................................................................... 219 Figure 4.8 Cadmium-block of extracellular calcium influx does not affect the frequency and the amplitude of mIPSCs in both WT and G203R mice. ....................................... 220 Figure 4.9 G203R mice showed a significant decrease in Gαo protein expression. ..... 222 Figure 4.10 G203R mice did not show any significant changes in Gβ expression in the brain. ............................................................................................................................. 223 Figure 4.11 Models of GABABR and α2AR mediated inhibition of GABA release. ........ 225 Figure 4.12 Gβγ may play a major role in the regulation of IPSCs. .............................. 227 Figure 4.13 GNAO1 mutations may have region specific effects, which cause an imbalance between excitatory and inhibitory neurotransmitters. .................................. 229 Figure S4.1 Despite the hypothesis that α2A receptor antagonist yohimbine (10µM) could reverse the inhibition of sIPSC frequency induced by UK14, 304, the application of yohimbine further reduced the sIPSC frequency with the application of UK14, 304. ... 231 Figure S4.2 Baclofen does not affect either frequency or amplitude of sIPSCs, and UK14,304 does not affect mIPSC frequency or amplitude. .......................................... 232 Figure S4.3 Heterozygous G184S (GOF) mice also showed a low Gαo protein expression level. ........................................................................................................... 233 Figure S4.4 Female Gnao1+/G184S mice showed reduced sIPSC frequency in hippocampal pyramidal cells, cortical layer II/IV pyramidal cells but not in cerebellar Purkinje cells. ................................................................................................................ 234 Figure S4.5 Brain Gαo expression does not change in R209H mice’s brain lysates. ... 235 Figure S5.1 α2AR activates Gαo, which inhibits N-type calcium channels in G1A1 cells. GOF mutation G184S enhance the inhibition of calcium currents. ............................... 273 Figure S5.2 Both G1A1 and SH-SY5Y cells are good candidates to study mutant Gαo’s xvi ! effects on N-type calcium channels with Fluo-4-NW dye in a Hamamatsu µCELL plate reader. .......................................................................................................................... 274 Figure S5.3 Neurite outgrowth in rat pheochromocytoma cells (PC12) can be induced by 50 ng/mL Nerve Growth Factor (NGF) in normal growth medium with reduced serum. ...................................................................................................................................... 275 Figure S5.4 Neurite outgrowth can be induced by 10 µM retinoid acid (RA) in normal growth medium with reduced serum (30% FBS) in human neuroblastoma cells (SH-SY5Y). ................................................................................................................... 276 Figure S5.5 Representative traces of modulations in calcium oscillations with different ion concentration in mixed cortical cultures from a WT mouse. ................................... 277 Figure S5.6 High-throughput assessment of neural excitability in cortical cultures. ..... 278 Figure S5.7 Nissl staining compares gross morphological changes in cerebellum from Gnao1+/G184S and Gnao1+/+ mice at age 8 weeks old. .................................................. 279 xvii ! KEY TO ABBREVIATIONS Adenylyl cyclases Artificial cerebral spinal fluid Attention Deficit Hyperactivity Disorder AC ACSF ADHD ANOVA Analysis of variance AP APV BRET DMEM CNQX DOR CREB CRISPR EA2 EIEE17 FBS FDR FFT Action potential 2-Amino-5-phosphonopentanoic acid Bioluminescence resonance energy transfer Dulbecco’s modified Eagle’s medium 6-Cyano-7-nitroquinoxaline-2,3-dione Delta opioid receptor cAMP-response element binding protein Clustered regularly interspaced short palindromic repeats Episodic ataxia type 2 Early infantile epileptic encephalopathy 17 Fetal bovine serum False discovery rate Fast Fourier Transform ! xviii GABA GDD GIRK GOF GPCR HD ID KO LID LOF MD M-D ME mEPSC mIPSC MOR NAc NEDIM γ-Aminobutyric acid Developmental delay G-protein activated inward rectifying potassium channels Gain-of-function G protein coupled receptor Huntington’s Disease Intellectual disability Knock-out L-DOPA-induced dyskinesia Loss-of-function Movement disorder Myoclonus-dystonia syndrome Mercaptoethanol Miniature excitatory postsynaptic current Miniature inhibitory postsynaptic current Mu opioid receptor Nucleus accumbens Neurodevelopmental disorder with involuntary movement xix ! NEM NF NGF PD PDE PI3-K PLOF PNKD3 PRRT2 PTX PTZ RNP sEPSC sIPSC SNAP25 ssODN SV TH N-Ethylmaleimide Normal-functioning Nerve Growth Factor Parkinson’s Disease Phosphodiesterase Phosphatidyl-inositol-4,5-bisphosphate 3-kinase Partial-loss-of-function Paroxysmal nonkinesigenic dyskinesia 3 Proline-rich transmembrane protein 2 Pertussis toxin Pentylenetetrazol Ribonucleoprotein Spontaneous excitatory postsynaptic current Spontaneous inhibitory postsynaptic current Synaptosomal associated protein-25 Single-stranded oligo DNA nucleotides Synaptic vesicles Tyrosine hydroxylase xx !! TTX WT Tetrodotoxin Wildtype xxi ! CHAPTER 1: AN OVERVIEW OF GNAO1-ASSOCIATED NEUROLOGICAL DISORDERS Modified from Feng, H., Khalil, S., Neubig, R. R., & Sidiropoulos, C. (2018). A mechanistic review on GNAO1-associated movement disorder. Neurobiology of disease, 116, 131-141. DOI: https://doi.org/10.1016/j.nbd.2018.05.005. With permission from the Elsevier. All rights reserved. 1 ! 1.1 Abstract Mutations in the GNAO1 gene cause a complex constellation of neurological disorders including epilepsy, developmental delay, and movement disorders. GNAO1 encodes Gαo, the α subunit of Go, a member of the Gi/o family of heterotrimeric G protein signal transducers. Go is the most abundant membrane protein in the mammalian central nervous system and plays major roles in synaptic neurotransmission and neurodevelopment. GNAO1 mutations were first reported in early infantile epileptic encephalopathy 17 (EIEE17), but are also associated with a more common syndrome termed neurodevelopmental disorder with involuntary movements (NEDIM). Here we review a mechanistic model in which loss-of-function (LOF) GNAO1 alleles cause epilepsy and gain-of-function (GOF) alleles are primarily associated with movement disorders. We also develop a signaling framework related to cyclic AMP (cAMP), synaptic vesicle release, and neural development and discuss gene mutations perturbing those mechanisms in a range of genetic movement disorders. Finally, we analyze clinical reports of patients carrying GNAO1 mutations with respect to their symptom onset and discuss pharmacological/surgical treatments in the context of our mechanistic model. 1.2 Introduction Mutations in GNAO1 were first reported in patients with Ohtahara syndrome and early infantile epileptic encephalopathy 17 (EIEE17, OMIM 61547) (Nakamura et al., 2 ! 2013). More recently, a syndrome of neurodevelopmental disorder with involuntary movements without epileptic seizures (NEDIM, OMIM 617493) has been defined, expanding the phenotypic spectrum of GNAO1 mutation-associated neurological disorders (Ananth et al., 2016; Zhu et al., 2015). Currently, there have been published reports on 81 patients representing 36 different GNAO1 mutations (23 missense, 1 in-frame deletion and 1 splicing site mutation, see Figure 1.1) (Ananth et al., 2016; Arya, Spaeth, Gilbert, Leach, & Holland, 2017; Blumkin et al., 2018; Bruun et al., 2018; Carecchio et al., 2019; Danti et al., 2017; R. Dhamija, Mink, Shah, & Goodkin, 2016; Dietel, 2016; Epi, 2016; Epi et al., 2013; Euro, Epilepsy Phenome/Genome, & Epi, 2014; Farwell et al., 2015; Gawlinski et al., 2016; Gerald et al., 2018; Helbig et al., 2016; Honey et al., 2018; Kelly et al., 2019; Koy et al., 2018; Kulkarni, Tang, Bhardwaj, Bernes, & Grebe, 2016; Law et al., 2015; Malaquias et al., 2019; Marce-Grau et al., 2016; Menke et al., 2016; Nakamura et al., 2013; Okumura et al., 2018; Saitsu et al., 2016; Sakamoto et al., 2017; Schirinzi et al., 2019; Schorling et al., 2017; Takezawa et al., 2018; Talvik et al., 2015; Ueda, Serajee, & Huq, 2016; Waak et al., 2018; Xiong et al., 2018; Yilmaz et al., 2016). Although recent reviews on monogenic complex hyperkinetic disorders recognized GNAO1 mutations as pathogenic (Carecchio & Mencacci, 2017; Mencacci & Carecchio, 2016), our review focuses on a mechanistic analysis illustrating the shared pathways of pathogenic mutations across multiple movement disorder-associated genes. It is 3 ! important to consider the mechanisms that underlie the GNAO1-associated movement disorders to rationalize the clinical heterogeneity resulting from different mutations in GNAO1, as well as the implications for therapeutic choices. We (H.F. and R.R.N.) recently demonstrated that GNAO1 mutations associated with movement disorders result in a gain-of-function (GOF) biochemical behavior related to control of cAMP levels, while epilepsy-associated mutations cause loss-of-function (LOF) behavior (Feng et al., 2017). This is consistent with other single-gene epilepsy and movement disorders, which also share causal genes (Batty, Fenrich, & Fouad, 2017; Szczepanik et al., 2015). Focusing on movement disorders, there is a clear functional connection between GNAO1 and other “movement disorder genes” related to two molecular mechanisms. Both the cAMP pathway (GNAL, GNB1, ADCY5, PDE10A) and regulation of synaptic vesicle fusion and neurotransmitter release (GNB1, CACNA1A, CACNA1B, KCNMA1, SYT1, SNAP25, and PRRT2) have been implicated. In this review, we attempt to develop models of these systems and explore how they may connect pathophysiology with clinical patterns and therapeutic responses. 4 ! Figure 1.1 Summary of reported cases of GNAO1 encephalopathy (A) Locations of reported mutations on the Gαo amino acid sequence. The splicing site mutations are not included here. (B) Sex distribution among the 48 patients reported. (C) Distribution of movement disorders and/or epilepsy symptoms in GNAO1 encephalopathy patients (Green = movement disorder only; Red = epilepsy only; Orange = both phenotypes). 1.3 Gαo (GNAO1) mechanisms GNAO1 encodes the α-subunit of a heterotrimeric guanine nucleotide-binding protein (Gαo), which is the most abundant membrane protein in the mammalian central nervous system, constituting approximately 1% of total brain membrane protein. Gαo localizes ubiquitously throughout the brain with relatively high expression in hippocampus, 5 ! striatum and cerebellum (Worley, Baraban, Van Dop, Neer, & Snyder, 1986). It couples to a variety of important G protein coupled receptors (GPCRs) including GABAB, α2 adrenergic, adenosine A1 (A1R), and dopamine D2 (D2R) receptors. These play key roles in regulating neurotransmitter release, movement, and neural development. There are multiple downstream signaling targets of Go, as well as of the other members of the Gi/o family. These include: inhibition of adenylyl cyclases (ACs) which decreases cAMP production, inhibition of N-type (Cav2.2) and P/Q type calcium channels (Cav2.1) (Colecraft, Brody, & Yue, 2001; McDavid & Currie, 2006), and direct inhibition of neurotransmitter vesicle release by the binding of Gβγ released from active Gαo to inhibit syntaxin 1A and SNAP25 (Zamponi & Currie, 2013). Both Gαo and Gβγ subunits also bind to G protein-coupled inward rectifying potassium (GIRK) channels to stimulate channel opening (Luscher & Slesinger, 2010). GIRK channels are well-recognized as playing a role in seizure disorders (Mayfield, Blednov, & Harris, 2015; Signorini, Liao, Duncan, Jan, & Stoffel, 1997; Torrecilla et al., 2002). Many of these targets of Go (both Gαo and Gβγ) signaling are also implicated in movement disorders. Mutations in ADCY5 (which encodes adenylyl cyclase type 5) have been reported in patients with dyskinesia and dystonia (Meijer, Miravite, Kopell, & Lubarr, 2017; Mencacci, Erro, et al., 2015; Shaw, Hisama, Friedman, & Bird, 1993). Mutations in CACNA1A (encoding Cav2.1) cause episodic ataxia type 2 (EA2) (Sintas et al., 2017; Wan et al., 2011). In the G protein family, mutations in both GNAL (Dufke et al., 2014; 6 ! Kumar et al., 2014; Putzel et al., 2016) and GNB1 (Lohmann et al., 2017; Steinrucke et al., 2016) are also associated with dystonic syndromes. The former encodes Gαolf which mediates dopamine D1/5 receptor stimulation of AC and the latter encodes Gβ1 which mediates many actions of Gi/o. 1.4 The clinical spectrum of GNAO1 mutation-associated movement disorders 1.4.1 GNAO1 encephalopathy displays a variety of neurological symptoms To understand the molecular mechanisms underlying GNAO1 disorders, it is important to consider the substantial clinical heterogeneity which includes both early-onset epileptic encephalopathy (Nakamura et al., 2013) and patients with complex movement disorders with or without epilepsy (Ananth et al., 2016; Kulkarni et al., 2016; Menke et al., 2016; Saitsu et al., 2016; Sakamoto et al., 2017; Zhu et al., 2015). Recently, we reported a biochemical analysis of 15 different GNAO1 mutant alleles (Feng et al., 2017), which revealed that LOF mutations are associated with epileptic seizures while mutations that result in GOF for inhibition of cAMP as well as mutations that show largely normal function in this assay (p.R209 mutations) are mainly associated with movement disorders (Feng et al., 2017). The two most common manifestations of patients with GNAO1 mutations (Table S1.1 and Figure S1), regardless of their clinical pattern or biochemical phenotype, are hypotonia (68%) and developmental delay (78%, Table S1.1). Choreoathetotic movements (44%) and dystonia (32%) are the next most common findings (Table S1.1). 7 ! Approximately 28% of patients had intellectual disability. While many individuals have abnormal EEG or MRI findings (Table S1.1), less than half of patients with GNAO1 mutations (50%) showed markedly abnormal EEGs and that was primarily in epilepsy patients with LOF mutations. Approximately 64% of the reported patients showed significant MRI findings and these were distributed across both epilepsy and movement disorder patients (Table S1.1). This heterogeneity in both clinical pattern and effect on brain structure/function suggests a role for both neurodevelopmental alterations and functional signaling perturbations. The latter seems more prominent in patients with the GOF mutants who show less evidence for brain structural abnormalities as well as having some therapeutic responses to drug treatment (see below). Table 1.1 Most Common GNAO1 Mutant Alleles Associated With Movement Disorders No. of Epileptic Chorea/ Frequency of occurring GNAO1 alleles patients seizures Hypotonia athetosis Dystonia Myoclonus Ballismus Dyskinesia Stereotypies p. R209H/L/G/C 12 p. E246K 9 25% 22% p. G203R 7 100% 83% 89% 14% 67% 63% 43% 33% 56% 29% p. E237K 2 0 100% 100% 100% 8% 0 0 50% 8% 50% 0 50% 25% 25% 0 50% 8% 0 0 0 ! 8 ! Table 1.2 Drugs showing beneficial effects to control involuntary movements of GNAO1 encephalopathy patients Drug Positive GNAO1 Response Mutations Sex Inheritance de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo p.E246G p.S47G p.R209H p.E237K Tetrabenazine p.E237K p.E246K p.E246K p.E246K p.E246K p.G45R p.E237K Levetiracetam Topiramate p.R209C Trihexyphenidyl p.R209H ! F M M M M F F M F M M F M Age of onset 6 mo 5 mo 10 mo 4 mo 3 mo 4 y 6 mo 14 y 3 mo NA 4 mo Symptoms Epileptic Movement seizures Disorders + + + + + + + + + + + + + + + + Ref Danti et al., 2017 Danti et al., 2017 Dhamija, 2016 This report Waak et al., 2017 Ananth et al., 2016 Ananth et al., 2016 Ananth et al., 2016 Waak et al., 2017 Ueda, 2016 This report Saitsu et al., 2016; Sakamoto et al., 2017 Dhamija, 2016 de novo 11 mo de novo 10 mo 1.4.2 Movement disorders related to GNAO1 encephalopathy show limited response to pharmacological treatments There are three mutation hotspots (G203, R209 and E246) in GNAO1 that prominently result in movement disorders (Table 1.1) (Ananth et al., 2016; Arya et al., 2017; Danti et al., 2017; R. Dhamija et al., 2016; Honey et al., 2018; Kulkarni et al., 2016; Menke et al., 2016; Nakamura et al., 2013; Saitsu et al., 2016; Sakamoto et al., 2017; Waak et al., 2018; Xiong et al., 2018; Zhu et al., 2015). All show GOF or normal function (NF) phenotypes in the in vitro cAMP regulation assay (Feng et al., 2017). This 9 ! correlation raises the possibility of rationalized drug selection in treating diseases related to GNAO1 mutations. Gαo-coupled-receptor antagonists should reduce signaling in movement disorder patients by decreasing the signal from hyperactive GOF GNAO1 mutants. Gαo-coupled-receptor agonists might be beneficial in epilepsy. In both cases, however, the receptors mediating the abnormal function would need to be identified. The NF mutant alleles (in p.R209) raise questions about the simple GOF/LOF model despite the fact that those patients show clear clinical movement disorder pathology. Moreover, the G203R mutation results in a modest GOF biochemical effect (Feng et al., 2017) but causes both movement disorder and frequent seizures – though the latter are reasonably easily controlled (Table S1.1). These mutations show that there is more to learn about the genotype-phenotype correlation. A key question will be whether another downstream signal (e.g. calcium and potassium channel regulation) may better correlate with clinical patterns. Among patients with movement disorders, tetrabenazine is the most effective drug (Table 1.2) (Ananth et al., 2016; Danti et al., 2017). This is not surprising given that tetrabenazine’s actions on VMAT2 will deplete multiple amine neurotransmitters (dopamine, norepinephrine, and serotonin). This should result in a wide-spread reduction of Go signaling through, for example, α2A adrenergic receptors, D2/D4 receptors, and 5-HT1 receptors. Responses were also reported to trihexyphenidyl, topiramate, and levetiracetam 10 ! (Table 1.2). Trihexyphenidyl is described as a selective muscarinic M1 receptor antagonist, but it binds to all five muscarinic receptors subtypes with similar affinity (Dorje et al., 1991). The M4 muscarinic receptor subtype is second most potently inhibited by trihexyphenidyl. That receptor is Gi/o coupled and strongly implicated in striatal function (Ztaou et al., 2016). This may be a therapeutic target worth serious consideration. Interestingly, the patient with the p.R209H mutation who responded to trihexyphenidyl also responded to tetrabenazine (R. Dhamija et al., 2016). Topiramate was very effective in suppressing chorea in a patient carrying a p.R209C mutation (Table 1.2) (Sakamoto et al., 2017). Levetiracetam also showed effectiveness in two patients. Neither of these latter two drugs is known to affect G protein coupled receptors. However, levetiracetam partly works by binding to its high-affinity binding site on a synaptic vesicle protein to inhibit neurotransmitter release globally (Grimminger et al., 2013; Ohno & Tokudome, 2017), which would explain its multi-functionality in suppressing both epilepsy and movement disorders. While multiple therapies have shown some efficacy in controlling involuntary movements, no drug seems to be able to mitigate developmental delay. 11 ! Table 1.3 Patients responding positively to deep brain stimulation (DBS) GNAO1 Age of Epileptic Movement Positive Mutation Sex Inheritance onset seizures Disorders Response Ref Symptoms DBS p.E246G F p.E237K M p.E246K F de novo de novo de novo 6 mo 3 mo 3 mo p.R209H M de novo 18 mo p.R209H M de novo 2 y F M F de novo 6 mo de novo 2 y de novo 13 mo p.R209C p,R209L ! p.Q233P + + + + + + + + + + Y Y Y Y Y Y Y Y Danti et al., 2017 Waak et al., 2017 Waak et al., 2017 Kulkarni et al., 2016 Kulkarni et al., 2016 Waak et al., 2017 Honey et al., 2018 Yilmaz et al., 2016 In contrast to the modest efficacy of drug treatment, in seven cases where DBS was performed, patients all responded well and involuntary movements were suppressed (Table 1.3) (Danti et al., 2017; Honey et al., 2018; Kulkarni et al., 2016; Waak et al., 2017; Yilmaz et al., 2016). Consequently, DBS targeting the globus pallidus pars interna (GPi), appears to be the most effective treatment for GNAO1 related movement disorders, at least in medication refractory cases. There are no reports to assess the effectiveness of DBS in other brain regions such as the subthalamic nucleus. DBS may be effective due to its general effects in modulating aberrant synchronization in the basal ganglia-thalamo-cortical loops (McIntyre & Anderson, 2016). There is no information yet on long-term-sustained efficacy however. Furthermore, there may be a publication bias 12 ! towards patients who responded well and DBS can have a prominent placebo effect, as seen for patients with Parkinson’s disease (de la Fuente-Fernandez, 2004; Mercado et al., 2006). Seizures in patients with GNAO1 mutations can be controlled to some degree by multiple anti-epileptic drugs (AEDs) (Danti et al., 2017). However, no drug has been shown to be particularly effective (Table S1.1). 1.5 Potential pathogenic mechanisms of GNAO1-associated movement disorders The possible etiological bases of GNAO1-associated movement disorder may be explained by examination of GNAO1 signaling. Inhibition of cAMP is a canonical pathway of Go, which may be mediated by Gαo itself or by the released Gβγ (Dortch-Carnes & Potter, 2003; Gill & Hammes, 2007). Mutations in ADCY5 (which encodes an AC protein that synthesizes cAMP) also result in movement abnormalities in human patients. Dysregulation of cAMP signaling leads to brain malfunction (Borlikova & Endo, 2009; Guan et al., 2011). Therefore disturbances of cAMP levels could disrupt a finely tuned neurodevelopmental system. A second theoretical basis of GNAO1-associated movement disorder relates to Go’s role in regulating neurotransmitter release. A close relationship has been proposed among neurotransmitter levels, brain morphology and behavioral experience (Goldstein, 2006). Deficiency of key neurotransmitters like catecholamines (dopamine, epinephrine and norepinephrine) and serotonin are widely studied in movement disorders or seizures 13 ! (Mercimek-Mahmutoglu et al., 2015). Go’s presynaptic role in regulating neurotransmitter release suggests another potential mechanism in the etiology of movement disorders. A third possibility, from a developmental view, could be alterations of neuronal maturation, which needs to occur at appropriate stages of neurological development. Therefore children with developmental defects would exhibit abnormal behaviors. It is striking that most patients with GNAO1-associated movement disorders also suffer from severe developmental delay. Morphologically, MRI scans may show global atrophy and delayed myelination. Overall, genetic causes are responsible for about 40% of the developmental delay cases, including developmental delay/intellectual disability (GDD/ID) (Miclea, Peca, Cuzmici, & Pop, 2015). Control of cAMP levels and neurotransmitter release clearly could affect ongoing neural functions as well as neurological development. By this concept, GNAO1-associated movement disorders could result from perturbations of either of these processes (Leung & Wong, 2017). Clearly, the former would be more amenable to the therapeutic intervention than the latter. 1.5.1 Role of cAMP regulation in movement disorders The second messenger cAMP modulates a broad spectrum of cellular functions including gene expression, metabolism, exocytosis, and neuronal development. cAMP is synthesized from ATP by the AC enzymes upon activation by G-protein coupled receptors (GPCRs). An increase of cAMP levels activates protein kinase A (PKA), which 14 ! phosphorylates other kinases, transcription factors, and ion channels. cAMP can also activate the Rap guanine nucleotide exchange factor Epac, which has important functions in neural plasticity (Schmidt, Dekker, & Maarsingh, 2013; Tong et al., 2017). cAMP formation is negatively regulated by phosphodiesterases (PDEs) or Gi/o family coupled GPCRs such as these associated with GNAO1. There are many isoforms of ACs and PDEs in specific brain regions, consistent with the need to maintain a delicate a balance of cAMP levels for a normally functioning nervous system. cAMP signaling in the brain is known to mediate neuronal excitability and synaptic plasticity, which further regulates learning and memory, and motor function (Bollen & Prickaerts, 2012; Kandel, 2012; Pierre, Eschenhagen, Geisslinger, & Scholich, 2009). Also, dibutyryl-cAMP, an analog of cAMP, promotes axon regeneration and the recovery of motor function by inhibiting the RhoA signaling pathway (Jeon et al., 2012). The importance of cAMP in pro-regenerative action makes it a potential therapeutic target for enhancing nerve repair (Yu, Wang, Wu, & Yi, 2017). In this section, we will focus on the role that cAMP plays in movement disorders. There are many movement disorder-related genes that directly regulate cAMP levels in the brain (D'Angelo et al., 2017; Padovan-Neto & West, 2017). Figure 1.2 provides a schematic overview of the genes discussed including the type of mutations (LOF or GOF) seen. However, see the text below for details, as not all results in the literature are clear-cut. Based on functional studies associated with those genes, both up- and down- 15 ! regulation of cAMP production may contribute to the pathophysiology of movement disorders. Here we provide a hypothetical model of the role of cAMP regulation in movement disorders (Figure 1.2). Unfortunately, however, there is not a simple, completely coherent model of the relationship between predicted changes in cAMP concentration and the presence of movement disorders. It is clear that perturbations in cAMP mechanisms are pathological. Potential explanations for this complexity are described below after consideration of each gene in detail. Figure 1.2 Genes regulating the cAMP pathway are related to movement disorders GPCRs activate Gαo (encoded by GNAO1), which may either inhibit or stimulate cAMP production depending on the AC subtype present. Gβ1 (encoded by GNB1) forms a complex with Gγ and this Gβγ complex, typically released from Go or Gi-family G proteins, can also inhibit or stimulate cAMP production. Activation of Golf (encoded by GNAL) stimulates cAMP. Phosphodiesterase 10A (encoded by the PDE10A gene) hydrolyzes 16 ! Figure 1.2 (cont’d) cAMP to the monophosphate. Both gain-of-function (GOF) mutations in GNAO1 and loss-of-function (LOF) mutations in GNAL likely result in a decrease of cAMP. LOF mutant alleles in PDE10A and GOF mutations in ADCY5 increase cAMP. 1.5.1.1 GNAL GNAL encodes the alpha subunit of the guanine nucleotide-binding protein Golf. It belongs to the Gs family and can be activated by odorant receptors in the olfactory epithelium. It is also strongly expressed in the striatum. Gs family and Gi/o family proteins have opposite functional effects on AC. Once activated by Gs-coupled GPCRs, Golf activates AC enzymes to produce cAMP. In the striatum, Golf couples to D1 dopamine (D1R) and A2A adenosine (A2AR) receptors to activate type 5 adenylyl cyclase (AC5), which is encoded by the ADCY5 gene (Mercimek-Mahmutoglu et al., 2015). GNAL mutations account for about 1% of all cases of focal or segmental dystonia (Kumar et al., 2014). These include autosomal dominant, partial LOF mutations in GNAL (Dos Santos et al., 2016; Masuho et al., 2016). As GOF mutations in GNAO1 and LOF mutations in GNAL result in a similar functional change in cAMP production, it seems logical that they may share similar mechanisms leading to dystonic/choreo-athetoid disorders. Other previous functional studies of mutant GNAL also revealed deficiencies in AC activation after D1R stimulation (Fuchs et al., 2013; Kumar et al., 2014). Typically, patients with heterozygous GNAL mutations exhibit an adult-onset focal cervical, laryngeal, and/or segmental dystonia (Masuho et al., 2016). Animal studies also support 17 ! the idea that LOF mutations in GNAL with a reduction in striatal cAMP lead to movement disorders. A heterozygous mouse model Gnal+/- was reported with abnormal postures and movements compared to WT mice after treatment with the muscarinic agonist oxotremorine (Pelosi, Menardy, Popa, Girault, & Herve, 2017) (Zwart, Reed, Clarke, & Sher, 2016) (Pelosi et al., 2017; Zwart et al., 2016). Note that oxotremorine-induced movement disorders in Gnal+/- mice can be replicated with infusion of oxotremorine into the striatum but not cerebellum (Pelosi et al., 2017), which indicates the crucial role of cAMP signaling in striatal projection neurons and the potential role of muscarinic receptors in modulating the motor movement. 1.5.1.2 GNB1 The GNB1 gene encodes the G protein β subunit Gβ1. In G protein signaling, Gα binds with Gβγ and GDP in its inactive state. Upon activation, Gα binds to GTP and the Gα-GTP and Gβγ separate, both carrying out downstream signaling. Recently, de novo mutations in GNB1 have been identified using whole-exome sequencing in patients with severe neurodevelopmental disability, hypotonia, and seizures (Lohmann et al., 2017; Petrovski et al., 2016). The symptoms in GNB1 patients share characteristics with GNAO1 encephalopathy patients. Moreover, patients with GNB1 mutations also display early onset of movement abnormalities similar to patients with GNAO1 mutations (Petrovski et al., 2016). However, the functional change of the known GNB1 mutations is unclear due a 18 ! variety of assays used by different groups. Lohmann et al defines their mutant GNB1 as LOF by using real-time bioluminescence resonance energy transfer (BRET) assays to assess Gβγ’s ability to couple to D1R (Lohmann et al., 2017). However, other groups describe GNB1 mutations as having a GOF effect due to enhancement of downstream signaling pathways (Petrovski et al., 2016; Yoda et al., 2015). One unified functional assay such as Gβγ-regulated inhibition of cAMP production or of N-type calcium channels should be performed on all GNB1 mutants to clarify definitions of LOF or GOF. Whether mutations in GNAO1 affect Gβγ function or not remains unknown, but the fact that non-functioning Cav2.1 and Cav2.2 result in similar movement disorders (see below) suggests a hypothesis that Gβγ inhibition of calcium channels may be enhanced by GOF mutations in GNAO1 as well as GOF mutations in GNB1 itself. 1.5.1.3 ADCY5 Mutations in the ADCY5 gene, which encodes AC5 also cause early onset persistent or paroxysmal choreic, myoclonic, and/or dystonic movements as well as alternating hemiplegia of childhood (Carapito et al., 2015; Friedman et al., 2016; Mencacci, Erro, et al., 2015; Westenberger et al., 2017). Patients carrying ADCY5 mutations display mixed hyperkinetic movements including dystonia, facial myokymia, chorea, myoclonus and tremor (D. H. Chen et al., 2015). In addition to abnormal movements, axial hypotonia with paroxysmal exacerbations is also associated with ADCY5 mutations (D. H. Chen et al., 2015). Delayed milestones and axial hypotonia seem to be almost universal features 19 ! in infants with ADCY5 mutations (Carecchio et al., 2017). This is very similar to patients with GNAO1 mutations (Table S1.1). One functional study measuring β-adrenergic agonist-stimulated intracellular cAMP shows that two de novo mutations (c.1252C>T, p.R418W and c.2176G>A, p.A726T) in ADCY5 are gain-of-function (GOF) mutations (Y. Z. Chen et al., 2014). More functional studies need to be done in assessing other mutations in ADCY5 to define a clear genotype-phenotype correlation. A GOF mutation in ADCY5 should increase cAMP levels contrasting with expected effects of GNAL and GNAO1 mutations. AC5 is highly expressed in striatum and nucleus accumbens (NAc). The striatum and NAc are part of the dopaminergic system that is activated in response to stress (Carapito et al., 2015). Chen et al reported that Adcy5-null mice developed a movement disorder, which can be worsened by stress (D. H. Chen et al., 2015). In addition, L-DOPA-induced dyskinesia (LID) is profoundly reduced in AC5 knockout mice and suppression of AC5 in the dorsal striatum is sufficient to attenuate LID (Park et al., 2014). Since Go inhibits AC, this is consistent with the fact that knockout AC5 animals show impaired movements (Iwamoto et al., 2003). However, it remains unclear why GOF mutations in ADCY5 lead to hyperkinetic movements in humans. These studies confirm the importance of regulation of cAMP in the development of movement disorders and suggest that inappropriate changes in either direction may be detrimental. 20 ! 1.5.1.4 PDE10A PDE10A (encoded by PDE10A gene) participates in signal transduction by regulating the levels of intracellular cyclic nucleotides. PDE10A is localized in dendritic spines proximally to postsynaptic sites in striatal medium spiny neurons (Xie et al., 2006). It hydrolyzes both cAMP and cGMP to nucleoside 5’ monophosphate (Russwurm, Koesling, & Russwurm, 2015). However, PDE10A is the major cellular mechanism for degradation for cAMP but has only modest activity for cGMP in the striatum (Russwurm et al., 2015). Recently, LOF mutations in PDE10A are found in infancy-onset hyperkinetic movement disorders and childhood-onset chorea (Diggle et al., 2016; Mencacci et al., 2016). Loss of striatal PDE10A associates with movement disorders like Hungtington’s and Parkinson’s disease (Ahmad et al., 2014; Giorgi et al., 2011). Pde10a knockout mice also show abnormalities in movements (Siuciak et al., 2008). LOF mutations in PDE10A increase cAMP concentration, which seems contradictory to GOF mutations in GNAO1. However, PDE10A levels differ in different striatum regions in an animal model of dystonia. PDE10A is increased in the globus pallidus but decreased in the entopeduncular nucleus/substantia nigra in a DYT1 model (D'Angelo et al., 2017), which lead to opposite regulation on cAMP concentration. This result further suggests the balance of cAMP concentration is more important in regulating neuronal functions than shifting either way. 21 ! 1.5.1.5 Summary of cAMP regulation in movement control cAMP has complex effects on neurotransmission. It can enhance neurotransmission through multiple actions on Ca++ concentration or on synaptic vesicle release and trafficking (Neher, 2006). It also strongly modulates synaptic plasticity – generally increasing neurotransmission (Nestler, Alreja, & Aghajanian, 1999), which may involve the cAMP-response element binding protein (CREB) or Epac (Eagle, Gajewski, & Robison, 2016; Tong et al., 2017). cAMP is also important in serotonin-mediated enhancement of synaptic transmission where it modulates hyperpolarization-activated cation channels (Ih channels), which underlie repetitive neuronal firing (Beaumont & Zucker, 2000). In addition to these relatively acute actions of cAMP on neurotransmission, it also plays key roles in neurodevelopment. cAMP activates neurite outgrowth and facilitates axonal guidance (Akiyama, Fukuda, Tojima, Nikolaev, & Kamiguchi, 2016; Inda et al., 2017). It is possible that gene mutations that elevate cAMP production (i.e. LOF mutations in PDE10A, GOF mutations in ADCY5) increase neurite outgrowth and may result in enhanced or disordered synapse formation. This may increase excitatory neurotransmission and neuron hyperexcitability. It is clear, however, that genetic mutations which lead to decreased cAMP production (i.e. LOF mutations in GNAL and GOF mutations in GNAO1) also result in movement disorders. Indeed for some movement-disorder-associated genes (i.e. GNB1), both LOF and GOF functions and 22 ! effects have been reported (Lohmann et al., 2017; Petrovski et al., 2016). There are several possible explanations for this highly complex picture. First, in vitro studies used to characterize GOF/LOF behavior of mutant alleles differ between labs so the concept of GOF or LOF may not directly correlate with in vivo cAMP production. Second, some actions may be mediated in different neuron populations or different brain regions. Finally, keeping an appropriate balance of cAMP levels may be the critical underlying element for normal movement performance. So either an increase or a decrease in cAMP could result in the observed movement disorders. 1.5.2 Role of neurotransmitter release and synaptic vesicle fusion in movement disorders In addition to the control of cAMP production, regulation of neurotransmitter release is also strongly implicated in movement disorders, epilepsy, and neurodevelopmental delays (Figure 1.3). Mutations in genes for a number of proteins directly involved in synaptic vesicle fusion (e.g. SYT1, SNAP25, and PRRT2) have been identified. Presynaptic calcium influx is another driving factor for neurotransmitter release and most CNS synapses rely on Cav2.1 (CACNA1A) or Cav2.2 (CACNA1B) calcium channels for synaptic transmission. Mutations in both of these channel genes are associated with movement disorders. Further tightening the connection among these various genes related to movement disorder, Gαo activation drives Gβγ release, which mediates direct inhibition of vesicle 23 ! release (Zurawski et al., 2017) as well as indirect inhibition by suppression of voltage-gated calcium channels (i.e. Cav2.1 and Cav2.2 encoded by CACNA1A and CACNA1B) (Agler et al., 2005; McDavid & Currie, 2006). These mechanisms generate a relatively consistent model in which reductions in neurotransmitter release associate with movement disorder. GOF mutations in both GNAO1 and GNB1 would result in increased activity of Gαo and Gβγ, which would suppress both calcium channel activity and the synaptic vesicle release machinery (Figure 1.3). LOF mutations in the genes encoding calcium channels (CACNA1A and CACNA1B) and three proteins involved in vesicle release (SYT1, SNAP25, and PRRT2) also are involved in movement abnormalities. It is well-established that calcium influx through CaV channels triggers synaptic vesicle fusion via synaptotagmin (Sudhof, 2012). Indeed presynaptic calcium channels form a complex with synaptotagmin and other proteins in the vesicle release machinery such as syntaxin 1 and SNAP-25 (Leveque et al., 1994; Simms & Zamponi, 2014). Also, Gβγ released from activated Go (Gαo/Gβγ) competes with synaptotagmin-1 for binding to SNARE proteins to modulate vesicle fusion (Zurawski et al., 2017), providing a tight network of mutant proteins controlling neurotransmitter vesicle release where suppressed synaptic vesicle release is associated with human movement disorders. 24 ! Figure 1.3 Pathogenic mutations in genes that regulate neurotransmitter release Activation of Gαo by multiple GPCRs (including GPR88) inhibits voltage gated calcium channels (CACNA1A and CACNA1B). Calcium influx promotes synaptotagmin-1 (encoded by SYT1) and SNAP25 (encoded by SNAP25 gene) anchoring the vesicle to the membrane in preparation for exocytosis. Reduced functions in CACNA1A, CACNA1B, GPR88, SYT1 and SNAP25 or increased function of GNAO1, GNB1 and KCNMA1 all reduce synaptic neurotransmitter releases. 1.5.2.1 SYT1 Synaptotagmin-1 (encoded by SYT1) is a calcium-binding synaptic vesicle protein required for both exocytosis and endocytosis. Tucker et al., 2004 investigated the effect of synaptotagmin-1 on membrane fusion mediated by the SNARE protein complex SNAP25, syntaxin and synaptobrevin. In the presence of calcium, the cytoplasmic domain of synaptotagmin-1 strongly stimulates membrane fusion (Figure 1.3). 25 ! Stimulation of fusion is abolished by disrupting the calcium-binding activity of synaptotagmin-1 (Rickman & Davletov, 2003). Thus, synaptotagmin-1 and SNAREs are likely to represent the minimal protein unit for calcium-triggered exocytosis (Tucker, Weber, & Chapman, 2004). The study of STY1’s relation to movement disorders is limited. Only one patient has been reported so far. A trio analysis of whole-exome sequences identified a de novo SYT1 missense variant (I368T) in a case of human neurodevelopmental disorder associated with hypotonia and hyperkinetic movements without seizures (Baker et al., 2015). Functional studies showed that this mutation slows evoked synaptic vesicle (SV) fusion and also affects SV retrieval from the plasma membrane during endocytosis (Baker et al., 2015). It is therefore a LOF mutation. Interestingly, an equivalent STY1 mutation in Drosophila also resulted in a reduction in evoked neurotransmitter release (Paddock et al., 2011). 1.5.2.2 PRRT2 Proline-rich transmembrane protein 2 (PRRT2) is encoded by the PRRT2 gene. Heterozygous mutations in PRRT2 lead to epilepsy, kinesigenic dyskinesia, and migraine. PRRT2 is enriched in presynaptic terminals. It regulates synapse number and release of SV. Most of pathogenic mutations in PRRT2 lead to impaired PRRT2 protein expression, which could result in impairment of neurotransmitter release (Weston, 2017). Moreover, PRRT2 protein interacts with the synaptic protein SNAP25 and 26 ! synaptotagmin-1 (H. Y. Lee et al., 2012; Weston, 2017). Prrt2 expression is high in mouse cerebral cortex, hippocampus, and cerebellum (W. J. Chen et al., 2011). Chen et al identified 3 heterozygous truncating mutations in PRRT2 gene from eight unrelated Han Chinese families with episodic kinesigenic dyskinesia-1 (2011). Independently, Wang et al identified one insertion and one nonsense mutation from 27 members of two families with autosomal dominant paroxysmal kinesigenic dyskinesias (Wang et al., 2011). Two patients in each family also developed an infantile convulsion and choreoathetosis syndrome (Wang et al., 2011). Law et al reported a common PRRT2 mutation in a case series of 16 patients with familial paroxysmal kinesigenic dyskinesia (Law et al., 2016). Heron et al identified heterozygous mutations in PRRT2 from separate families with familial infantile seizures-2 and with familial infantile convulsions with paroxysmal choreoathetosis (Heron et al., 2012), suggesting that mutations in PRRT2 are pathogenic for both epilepsy and movement disorders. 1.5.2.3 SNAP25 Synaptosomal associated protein-25 (SNAP25 encoded by SNAP25) is a component of the SNARE complex, which is essential to synaptic vesicle exocytosis. It also negatively modulates neuronal voltage-gated calcium channels by directly interacting with calcium channel subunits. The SNAP25 gene is associated Attention Deficit Hyperactivity Disorder (ADHD), schizophrenia, and bipolar disorder (Antonucci et al., 2016; Corradini, Verderio, Sala, Wilson, & Matteoli, 2009). In 2013, whole exome 27 ! sequencing identified a novel de novo mutation p.F48V (c.142G>T) in SNAP25 from a 15y old female with severe static encephalopathy, intellectual disability and generalized epilepsy (Rohena et al., 2013). SNAP25 also plays a major role in neuronal survival. Neuronal cultures from Snap-25 knockout mice show degenerated dendrites and ultimately neuronal death (Delgado-Martinez, Nehring, & Sorensen, 2007). The relationship between SNAP25 and Go has been well-studied. SNAP25 is a key downstream target of Gβγ subunits. Gβγ binds to the extreme C terminus of SNAP25 to inhibit vesicle release in response to Gi/o-coupled receptors activation (Zurawski, Rodriguez, Hyde, Alford, & Hamm, 2016). 1.5.2.4 KCNMA1 KCNMA1 encodes the pore-forming subunit of calcium-activated potassium channels (BK channels), which are in close proximity with voltage-gated calcium channels in neurons. Membrane depolarization activates calcium channels and increases calcium entry, which activates BK channels to help terminate the action potential, to produce after hyperpolarization, and to block calcium channels (U. S. Lee & Cui, 2010). Both GOF and LOF mutations of KCNMA1 were reported in patients with paroxysmal nonkinesigenic dyskinesia 3 (PNKD3), with or with out generalized epilepsy. This highlights the sensitivity of developing brain to both increased and decreased BK channel activities. Similar to GNAO1-associated movement disorders, developmental delay is also commonly associated with PNKD3 patients (Tabarki, AlMajhad, AlHashem, 28 ! Shaheen, & Alkuraya, 2016; Yesil et al., 2018; Zhang, Tian, Gao, Jiang, & Wu, 2015). GOF mutants of GNAO1 may work to regulate the function of BK channels both positively and negatively. Go activates the production of phosphatidyl- inositol-4,5-bisphosphate 3-kinase (PI3-K), which in turn activates BK channels (Patel, 2004; Shanley, O'Malley, Irving, Ashford, & Harvey, 2002). However, the Gβγ subunit dissociated from Gαo inhibits voltage-gated calcium channels (illustrated in 3.2.5 and 3.2.6), which would reduce intracellular calcium concentrations preventing the activation of BK channels (Castillo et al., 2015). Therefore, it is difficult to pinpoint the relationship between GOF mutations in GNAO1 and mutations in KCNMA1. 1.5.2.5 CACNA1A CACNA1A encodes the P/Q type voltage gated calcium channel α1 subunit (CaV2.1). Similar to N-type calcium channels (CaV2.2), activation of P/Q type calcium channels promotes neurotransmitter release. Moreover, influx of calcium through P/Q-type channels is responsible for activating expression of syntaxin-1A, a presynaptic protein that mediates vesicle docking (Sutton, McRory, Guthrie, Murphy, & Snutch, 1999). LOF mutations in CACNA1A also cause episodic ataxia type 2 (EA2), an autosomal dominant neurological disease (Guida et al., 2001; Jen, Yue, Karrim, Nelson, & Baloh, 1998; Sintas et al., 2017; Wan et al., 2011) and familial hemiplegic migraine type 1 (Garza-Lopez et al., 2012; Mullner, Broos, van den Maagdenberg, & Striessnig, 2004). There is evidence of a dominant negative effect of EA2 mutants in the CACNA1A gene 29 ! (Gao et al., 2012; Jeng, Chen, Chen, & Tang, 2006; Jouvenceau et al., 2001). Haploinsufficiency is also pathologic (Guida et al., 2001; Wan et al., 2011). The leaner mutation in mice affects the P/Q-type calcium channels Cav2.1 subunits, causing a reduction in calcium currents, predominantly in cerebellar Purkinje cells (Alonso et al., 2008). Homozygous leaner mice show severe, progressive cerebellar ataxia from postnatal day 10 (Alonso et al., 2008). Age-dependent impairment in motor and cognitive tasks is also observed in heterozygous leaner mice (Alonso et al., 2008). In addition, silencing of P/Q-type calcium channels in Purkinje neurons of adult mouse leads to a phenotype similar to episodic ataxia type 2 (EA2) (Salvi et al., 2014). Thus P/Q-type calcium channels play an important role control of movement as well as in neurodevelopment. 1.5.2.6 CACNA1B A disruptive missense mutation p.R1389H (c.4166G>A) in the CACNA1B gene, encoding neuronal voltage-gated N-type calcium channels (Cav2.2), was identified in a new familial myoclonus-dystonia (M-D) syndrome (J. L. Groen et al., 2015). Five affected family members were identified in a 16-member family across 3 generations (J. Groen, van Rootselaar, van der Salm, Bloem, & Tijssen, 2011). But a genome-wide study in a large European multicentric M-D cohort failed to detect the mutation in the 146 probands with familial M-D (Mencacci, R'Bibo, et al., 2015). Therefore, a causal association between the CACNA1B mutation p.R1389H (c.4166G>A) and movement disorder is still 30 ! under debate. However, a recent study linked LOF mutations in CACNA1B to the onset of neurodevelopmental disorder with seizures and nonepileptic hyperkinetic movements (NEDNEH; OMIM# 618497), which provides evidence for the role of Cav2.2 in human neurodevelopment (Gorman et al., 2019). 1.5.2.7 GPR88 GPR88 (encoded by GPR88 gene) is an orphan G-protein coupled receptor (GPCR) in the rhodopsin-like receptor family. It is widely expressed in the striatum, caudate nucleus, putamen, nucleus accumbens, and olfactory tubercle, but is not detected in the cerebellum (Massart, Guilloux, Mignon, Sokoloff, & Diaz, 2009). Its CNS expression is particularly robust in the striatum, paralleling that of the dopamine D2 receptor (Mizushima et al., 2000). Striatal GRP88 is enriched in both D1 and D2 expressing medium spiny neurons (Jin et al., 2014) and it is emerging as a key player in the pathophysiology of several neurological diseases. Agonists of GPR88 were developed as potential treatment for CNS disorders such as schizophrenia (Bi et al., 2015). One case of GPR88-associated chorea has been reported in human patients. Alkufri et al reported a deleterious mutation p.C291X in GPR88 associated with chorea, speech delay and learning disabilities (Alkufri, Shaag, Abu-Libdeh, & Elpeleg, 2016). Homozygous Gpr88 knockout mice displayed reduced striatal dependent behaviors such as rearing, grooming, and burying (Meirsman et al., 2016). Mechanistic studies in striatal medium spiny neurons demonstrated increased glutamatergic responses and reduced 31 ! GABAergic inhibition in the absence of GPR88 which results in enhanced neuronal firing in vivo (Quintana et al., 2012). Studies also showed enhanced function of Gi/o-coupled delta and mu opioid (DOR and MOR) in striatal membranes in Gpr88 knockout mice, suggesting a functional antagonism between GPR88 and other Gi/o-coupled receptor activities (Figure 1.3) (Meirsman et al., 2016). The increasing interest in GPR88 may result in the design of treatments for this orphan disease, considering its unique location. 1.5.2.8 Summary of neurotransmitter release and synaptic vesicle fusion in movement control The disruption of neurotransmitter release leads to a broad spectrum of movement impairments including chorea, ataxia and dystonia but also results in neurodevelopmental abnormalities and epilepsy. Apart from the relatively constrained expression of GPR88 in striatum, the rest of the proteins mentioned above are ubiquitously expressed throughout the brain. Therefore, it is relatively hard to determine which brain regions are involved in the pathogenesis of the related movement disorders. Hence, clear delineation of the neural mechanisms and relevant pathways remains challenging. The fact that GNAO1 is involved in the regulation of many of the proteins encoded by the above-mentioned genes makes Go-coupled receptors a possible therapeutic target in developing treatments for patients carrying these mutations. 1.6 Developmental defects may also contribute to movement disorders In addition to ongoing alterations in cAMP signaling and neurotransmitter release 32 ! mechanisms, it is likely that developmental abnormality that results in disordered synaptic or neural pathway organization could contribute to the observed movement disorders. It is estimated that genetic factors are responsible for up to 40% of developmental disabilities (Miclea et al., 2015). Such changes in neural development and organization could contribute to the observation that developmental delay and epilepsy are often seen with the gene mutations discussed in this review. However, the development effects and the physiological effects may not be mutually exclusive. cAMP signaling affects many developmental processes including facilitation of neuronal cell differentiation and maturation (Lepski, Jannes, Nikkhah, & Bischofberger, 2013; Sharma, Hansen, & Notter, 1990), induction of axon growth (Corredor et al., 2012), control of guidance cue (Forbes, Thompson, Yuan, & Goodhill, 2012), and enhancement of synaptic connections (Lessmann & Heumann, 1997). This could explain why most patients carrying the GOF mutations in GNAO1, which would be expected to suppress cAMP levels, exhibit severe developmental delay. Similarly, synaptic release mechanisms are implicated in both developmental abnormalities and movement disorders. Specifically, the individual carrying the LOF mutation in SYT1 with hypotonia and hyperkinetic movements discussed above also exhibited severe motor delay and profound cognitive impairment (Baker et al., 2015). Clearly, therapeutic approaches to modulate ongoing cAMP signaling or synaptic neurotransmitter release will be more tractable than attempting to alter developmental 33 ! abnormalities that may have already occurred by the time that the disorder is recognized. 1.7 Conclusion A full understanding of the exact etiology of GNAO1-related movement disorders remains elusive, however, the mechanistic clustering of GNAO1 and other movement disorder-related genes in the pathways for control of cAMP and neurotransmitter release strongly suggest a role for those mechanisms. For synaptic vesicle release, it appears that loss of function is the primary alteration observed. In contrast, for cAMP control, the picture is more confusing with mutations predicted to cause either increases or decreases being implicated. Clearly more needs to be learned about the specific brain regions, neuron types, and receptors that control GNAO1 signaling in these disorders. This may be facilitated by the development of new animal models beyond the two GNAO1 mutants already reported (Jiang & Bajpayee, 2009; Jiang et al., 1998; Kehrl et al., 2014). In particular, it is surprising that Gnao1+/- mice, which have relatively normal behaviors, do not mimic patients with heterozygous LOF GNAO1 mutations who exhibit severe epileptic encephalopathy. Animal knock-in models carrying the specific GNAO1 mutant alleles that are found in human epilepsy and movement disorders may also help clarify this conundrum. In addition to the mechanistic analysis, our review of the clinical literature also has implications for therapeutics. Since GOF mutations in GNAO1 cause movement disorders (Feng et al., 2017), it is not surprising that tetrabenazine has proven to be the 34 ! most effective drug in controlling patient’s chorea (Table 1.3) (Ananth et al., 2016; Danti et al., 2017; R. M. Dhamija, J, W.; Shah, B, B.; Goodkin, H, P., 2016; Waak et al., 2017). The broad depletion of multiple monoamines such as dopamine, serotonin, norepinephrine, and histamines from nerve terminals suggests that many receptors may be involved. A better understanding of the associated upstream GPCRs that are driving the enhanced signaling through Go may permit the use of more selective agonists or antagonists against these receptors to alleviate symptoms with fewer side effects in GNAO1 associated movement disorders. 1.8 Organization of the thesis This dissertation shows a genotype-phenotype correlation model of the GNAO1-associated neurological disorders and how this model helps further mechanistic study of the GNAO1-related neurological disorders. The organization of this thesis is as follows. Chapter 1 gives a broad review of GNAO1 encephalopathies and provides analysis of the pathophysiological mechanisms. The appendix includes a review of each case of reported human patients carrying GNAO1 mutations. Chapter 2 introduces the genotype-phenotype correlation model based on in vitro biochemical functional analysis. The appendices verify and expand this model with newly reported GNAO1 mutations. Chapter 3 provides a detailed behavioral description on the animal models carrying the human mutations G203R, R209H and T191F197del respectively. These three animal models are compared with the previously reported mouse model carrying a GOF Gnao1 35 ! mutant G184S and the Gnao1 KO mice. The appendix includes all the raw data collected from the behavioral experiments. Chapter 4 explores the electrophysiological mechanisms of the cerebellar Purkinje cells of the G203R mutant mice. The appendix expands the analysis to other mouse models mentioned in Chapter 3. Chapter 5 summarizes and analyzes all the experimental data collected from Chapter 2 to Chapter 4. In addition, this chapter speculates on the future directions and the significance of this project. This chapter also shows preliminary results of experiments that have not been developed maturely in the appendix. These results may help the future efforts of optimizing assays for further testing of GNAO1 mutations and for drug development or repurposing. 36 ! APPENDIX 37 ! APPENDIX SUPPLEMENTAL DATA Table S1.1 A complete summary of clinical information regarding GNAO1 patients No. GNAO1 Mutations Sex Inheritance Age of onset Seizures Hypotonia Chorea/athetosis Dystonia Myoclonus Ballismus Dyskinesia Stereotypies Developmental Delay Intellectual Disability Other Reported involvement F F M M F M F F M F F F F F F de novo de novo de novo de novo birth 3 y birth 2.5 mo de novo 5 wk de novo 15 hr de novo 2 hr de novo infancy de novo NA de novo, somatic mosaic de novo 29 d 14 d de novo 6 mo de novo de novo 3 mo 2 mo de novo 2.5 mo M de novo 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 p.G40R c.118G>A p.G40R c.118G>A p.G40R c.118G>A p.G40R c.118G>A p.G40W c.118G>T p.G40E c.119G>A p.G40E c.119G>A p.G45E c.134G>A p.G45R c. 133G>C p. D174G c.521A>G p.T191_F197del c.572_592del p.R349_G352del insQGCA c.1046_1055del1 0ins10 p.L199P c.596T>C p.A227V c.680C>T p.Y231C c.692 A>G p.Y231C c.692A>G p.E246G c.737A>G p.N270H c.808A>G p.D273V c.818A>T p.D273V c.818A>T p.F275S c.824T>C p.I279N c.836T>A p.I279N c.836T>A p.I279N c.836T>A p.Y291N c.871T>A p.S47G c.139A>G p.I56T c.167T>C p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G203R c.607G>A p.G204R c.610G>C p.G42R c.124G>C p.S207Y c.620C>A F F F F F M F M F M F F F F M F F F F M M F M 5 d 6 mo 3 mo 2 d 2 d 3 d 9 d 4 d 1 h 2 mo 5 mo 4 y 7 mo 7 d 9 d 1 mo 3 mo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo de novo birth de novo de novo de novo birth birth 12 d de novo 24 mo de novo de novo NA infancy + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + 38 ! + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + microcephaly ataxia + + + + + PEHO Syndrome; Accompanying mutation HESX1 (A9T) cerebellar ataxia, accompanying with ATP2B3 (T113M) Ohtahara syndrome Ohtahara syndrome + oromotor apraxia tetraparesis acquired microcephaly Ohtahara syndrome lower limb spasticity Ohtahara syndrome bradycardia + + + + + + + + + + + tachycardia, hyperthermia, sweating frequent arching of the back Table S1.1 (cont’d) No. EEG Findings Severe EEG a disorganized background with frequent multi-focal high- amplitude sharp and spike wave discharges, no definite burst-suppression patterns burst suppression at onset; slow background multifocal spike waves discontinous background, bilateral yemporal sharp waves 3 y: slow spike and wave, multifocal spikes 9 mo: right temporal seizures, focal spikes, focal slowing 2 mo: hypsar- rhythmia; 14 y: generalized onset of tonic seizures and epileptic spasms, generalized slowing 9 d: multifocal spikes; 14 y: focal spikes and waves, absence of normal awake and sleep features NA NA burst suppression at 2 mo; hypsarrhythmia at 3 mo; diffuse spike-and-slow-wave complex at 1yr 7 mo; sharp waves at frontal lobe at 3 yr 9 mo suppression-burst pattern at 2 week; hypsarrhythmia at 4 mo activated posterior temporal and occipital spikes 7 y: sleep! background slowing, multifocal high-voltaged sharp waves and spike and slow-wave complexes hypsarrhythmia and multifocal with ictal modified burst-suppression at 3 mo; hypsarrhythmia in awake burst-suppression pattern in sleep at 7 mo; slowing of background activity with multifocal interictal epileptiform discharge at 9 mo Neonatal: multifocal epileptiform sharp waves; 20 mo: frequent bioccipital spikes normal hypsarrhythmia and slow background Abnormal epileptiform activity normal at first; later unkown burst-suppression pattern, hypsarrhythmia, later slow background with multifocal discharges multifocal modified hypsarrhythmia, burst-suppression in sleep, high-voltage midline central discharges during spasms, generalized decrement with low voltage fast activity burst-suppression pattern at 4 d; multifocal sharp waves at 1 yr, 4 mo; burst-suppression pattern ar 5 yr, 6 mo Neonatal: multifocal sharp waves with high! amplitude bursts (not burst suppression); 8 mo: modified hypsarrhythmia 1 y: multifocal spikes, focal seizures; 3 y: intermittent posterior slowing right frontotemporal spikes left frontotemporal spikes diffuse irregular spike-and-slow-wave complex at 5 yr slow-wave bursts, migrating focal epileptiform discharges delta and theta activity and rare multi-regional, bi- hemispheric epileptic activity multifocal and diffuse discharges, along with generalized- onset seizures background slowing multifocal sharp waves, left temporal seizure pattern hypsarrhythmia multifocal paroxysmal activities in both temporal hemispheres NA NA NA 15 mo: slow posterior dominant rhythm 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ ++ + + ++ ++ + ++ ++ ++ ++ Brain MRI normal Severe MRI mild ventricular enlargement; thin corpus callosum bilateral increased signal in frontal and peritrigonal white matter in T1 4 mo: mildly prominent bifrontal subarachnoid spaces 4 mo: bilateral mesial temporal sclerosis, diffuse parenchymal atrophy, delayed myelination 2 y: status posttemporal lobectomy, left cerebral atrophy 15 mo: nonspecific signal increase in globi pallidi, normal myelination cerebral and cerebellar atrophy NA delayed myelination and thin corpus callosum at 10 mo normal at 3 mo normal delayed myelination and thin corpus callosum progressive cerebral atrophy, thin corpus callosum at 10 mo delayed myelination, short and thin corpus callosum and hippocampus 2 y: prominent subarachnoid spaces mild loss of volume (atrophy) in generalized distribution minimal atrophy normal normal at first; later unknown delayed myelination and thinning of white matter moderate-severe progressive global atrophy with delayed myelination, thin corpus callosum normal at 1 mo; cerebral atrophy at 5 yr 6 mo 2.5 y: moderate to progressive atrophy with delayed myelination normal ventricular enlargement; thin and dysmorphic corpus callosum; mild hypoplasia of caudate left frontal lesion (diffuse astrocytoma WHO grade 2); low-lying cerebellar tonsils delayed myelination at 1 yr, 3 mo; reduced cerebral white matter, thin corpus callosum at 4 yr, 8 mo progressive cerebral atrophy with delayed myelination at 14 mo mild atrophy progressive diffuse cerebral atrophy and volume loss in cerebellum atrophy, thin corpus callosum (2 y) mild atrophy (10 mo) NA thin corpus callosum hypomyelination and atrophy 14 y: bilateral hyperintensities of the thalamus on T2 NA 1 y: generalized thinning of corpus callosum, relative paucity of deep white matter 39 ! + + + ++ ++ ++ ++ ++ ++ ++ ++ + + + ++ ++ ++ ++ ++ ++ ++ ++ + ++ + + NA + + + ++ Table S1.1 (cont’d) No. Drug Trialed Drug Positive Response DBS DBS Positive Response (cid:7)(cid:11)(cid:12)(cid:8)(cid:13) (cid:6)(cid:4) (cid:2)(cid:1)(cid:9)(cid:10)(cid:1)(cid:2)(cid:5) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 NA multiple AEDs, ketogenic diet, vagus nerve stimulation NA NA NA NA NA NA NA NA NA NA NA NA NA NA pyridoxine, phenobarbital, levetiracetam, topiramate, vigabatrin, ACTH, zonisamide, clobazam ACTH, levetiracetam, clobazam, zonisamide NA NA NA ketogenic diet NA NA NA NA prednisone, valproic acid, clonazepam prednisone phenobarbitone, vigabatrin, lamotrigine NA NA NA NA N NA tetrabenazine NA NA lamotrigine, valproic acid, trihexyphenidyl and melatonin seisures and dyskinetic movements well-controlled NA valproate, ketogenic diet, pyridoxal-5-phosphate, vigabatrin, levetiracetam, phenobarbitone, prednisolone NA NA NA NA NA NA NA NA NA NA NA tetrabenazine, well-controlled on AED well-controlled on AED NA phenobarbital phenobarbital lamotrigine, zonisamide NA Vit B6, sulthiame, levetiracetam, L-DOPA Vit B6, phenobarbital, levetiracetam, topiramate, valproic acid, vigabatrin, oxcarbazepine, phenytoin, clobazam, lacosamide, ketogenic diet, gabapentin, trihexyphenidyl, baclofen, benzodiazepines controlled epileptic activity but not involuntary movements NA N lamotrigine and zonisamide controlled epileptic activity, benzodiazepines reduced paroxysmal dystonias NA topiramate, vigabatrin phenobarbital, carbamazepine, benzodiazepines, topiramate, clonazepam topiramate, clonazepam tetrabenazine, lorazepam, baclofen and phenobarbital phenobarbital N N N N N N N N N N N N N N N N Y N N N N N N N N N N N N N N N N N N N NA NA NA tetrabenazine Evaluated NA NA N N 40 ! N N N N N N N N N N N N N N N Law et al 2015 Danti et al 2017 Bruun et al 2017 Kelly et al 2019 Kelly et al 2019 Kelly et al 2019 Kelly et al 2019 Gawlinski et al 2016 26485252 28357411 28817111 30682224 30682224 30682224 30682224 27343026 Ueda et al 2015 DOI: 10.1055/s-0036-1597627 Nakamura et al 2013 Nakamura et al 2013 Kelly et al 2019 Marce-Grau A et al 2016 Saitsu et al 2015 23993195 23993195 30682224 27072799 25966631 Talvik et al 2015 DOI: 10.1177/2329048X15583717 NA Kelly et al 2019 Danti et al 2017 EuroEPINOMICS-RES Consortium 2014 Kelly et al 2019 Schirinzi et al 2018 EuroEPINOMICS-RES Consortium 2014 Epi4k 2016 Nakamura et al 2013 Kelly et al 2019 Kelly et al 2019 Danti et al 2017 Danti et al 2017 Nakamura et al 2013 Saitsu et al 2015 Y N N N N N N NA NA N N N N N N N N N N N N N 30682224 28357411 25262651 30682224 30642806 25262651 27476654 23993195 30682224 30682224 28357411 28357411 23993195 25966631 29429466 30642806 30642806 30103967 25590979 30682224 Dietel et al 2016 DOI: 10.1055/s-0036-1583625 Arya et al 2017 Schorling et al 2017 28202424 28628939 Schorling et al 2017 28628939 Xiong et al 2018 Schirinzi et al 2018 Schirinzi et al 2018 Koy et al 2018 Zhu et al 2014 NA Kelly et al 2019 Table S1.1 (cont’d) (cid:2)(cid:3)(cid:1) 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61# 62 63 64 65 66 67 68 69 70 71 72* 73* 74 75 76 77 78 79 80 81 GNAO1 Mutations p.R209H c.626G>A p.R209H c.626G>A p.R209H c.626G>A p.R209H c.626G>A p.R209H c.626G>A p.R209H c.626G>A p.R209H c.626G>A p.R209L c.626G>T p.R209G c.625C>G p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p.R209C c.625C>T p. C215Y c. 644G>A p.A221D c.662C>A p.Q233P c.698A>C p.E237K c.709G>A p.E237K c.709G>A p.E237K c.709G>A p.E237K c.709G>A p.E237K c.709G>A p.E237K c.709G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.E246K c.736G>A p.L284S c.851T>C p.I344del c.1030_1032delA TT c.723+1G>T c.723+1G>A Sex Inheritance Age of onset Seizures Hypotonia Chorea/athetosis Dystonia Myoclonus Ballismus Dyskinesia Stereotypies Developmental Delay Intellectual Disability Other Reported involvement M M M M M F M M F F M M F F F F F M M F F M M F F M M F M F F M M F F F F F F F F F de novo de novo de novo de novo de novo de novo de novo de novo 1 y 18 mo 2 y 10 mo 3 y 6 mo 6 mo birth de novo 3 y 10 mo de novo 11 mo de novo de novo birth 7 mo de novo 6 mo de novo de novo de novo 6 mo 6 mo birth de novo neonate + + + de novo infant + de novo de novo 12 yr 9 mo de novo 13 mo de novo de novo de novo de novo de novo de novo de novo de novo de novo 4 mo 3 mo 6 mo 6 mo neonate neonate 4 mo 4 y 4 y de novo 6 mo de novo de novo de novo de novo 14 y 11 mo 3 mo 3 mo de novo 13 mo de novo de novo 4 yr 30 mo de novo 11 d de novo 12 mo de novo 3 y de novo 4 mo + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + 41 ! + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + excessive movements tarchycardia ataxia dysarthria quadriplegia excessive movements; microcephaly gait ataxia peripheral spasticity tachycardia, hypertension microcephaly single seizure at 4 yr tachycardia, hyperthermia, sweating microcephaly + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + Table S1.1 (cont’d) No. EEG Findings Severe EEG Brain MRI Severe MRI 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61# 62 63 64 65 66 67 68 69 70 71 72* 73* 74 75 76 77 78 79 80 81 normal no irregularities other than diffuse slowing no irregularities other than diffuse slowing NA NA normal NA normal NA diffused low activities low background activities bilateral centrotemporal spikes normal diffuse slow activity 15 mo: slow posterior dominant rhythm normal NA NA NA 11 y: normal; 15 y: abnormal during sleep, frequent sharp waves NA NA NA NA NA NA NA normal at 12 yr NA NA NA NA normal right-sided polyspike-wave formations NA NA NA NA suppression-burst pattern normal NA NA + + + + + ++ ++ 42 ! normal normal normal normal global atrophy at 15 yr 13 mo: frontal lobe volume loss normal normal normal at 13 mo progressive cerebral and cerebellar atrophy, brainstem atrophy, thin corpus callosum ventricular enlargement; thin corpus callosum; mild hypoplasia of caudate; hypoplasia of inferior vermis normal normal posterior left periventricular hypersignal 1 y: generalized thinning of corpus callosum, relative paucity of deep white matte temporal atrophy, ventricular enlargement and mild temporal hypomyelination Cortical (prominent frontally) subcortical atrophy. Bilateral hypointense signals of globus pallidus on SWI and T2* MRI sequences Cortical (prominent frontally) subcortical atrophy. Bilateral hypointense signals of globus pallidus on SWI and T2* MRI normal normal NA NA progressive global atrophy (8 y, 12 y) NA 13 mo: mild hyperintensity in the occipital white matter Small medio- putaminal atrophy normal normal at 4 and 12 yr normal at 12 mo global atrophy at 5.5 yr global atrophy and T2 hypointensity in globus pallidi at 9 yr T2 hypointensity in globus pallidi at 14 yr normal atrophy of right hippocampus progressive global atrophy (1y, 5y, 8y) NA mild diffuse cortical atrophy slight hyperintensity of the left pars triangularis diffuse cerebral atrophy normal 8 yr: cerebral atrophy ventricular enlargement and dilated subarachnoid spaces; moderate cortical atrophy; dysmorphic corpus callosum; mild hypoplasia of caudate muclei ++ ++ ++ ++ + ++ ++ ++ ++ ++ + + ++ ++ + + ++ + ++ ++ ++ Table S1.1 (cont’d) No. Drug Trialed Drug Positive Response DBS NA N N NA Y Y tetrabenazine, trihexyphenidyl Evaluated Study PMID or DOI Menke et al 2016 Kulkarni et al 2015 Kulkarni et al 2015 27625011 26060304 26060304 Dhamija et al 2016 DOI: 10.1002/mdc3.12344 DBS Positive Response NA Y Y NA N N N NA Menke et al 2016 Ananth et al 2016 Kelly et al 2019 Blumkin et al 2018 Ananth et al 2016 Saitsu et al 2016; Sakamoto et al 2017 Danti 2017 Danti 2017 Waak et al 2018 Malaquias et al 2019 Kelly et al 2019 Schirinzi et al 2018 27068059 30682224 29801190 27625011 27068059 25966631; 27916449 28357411 28357411 28668776 31190250 30682224 30642806 Koy et al 2018 30103967 Koy et al 2018 Carecchio et al 2019 Kelly et al 2019 Yilmaz et al 2016 Feng at al. 2018 Waak et al 2018 Okumura et al 2018 Schirinzi et al 2018 Koy et al 2018 Koy et al 2018 Saitsu et al 2015 Ananth et al 2016 Ananth et al 2016 Ananth et al 2016 Ananth et al 2016 Schorling et al 2017 Schorling et al 2017 Waak et al 2018 Carecchio et al 2019 Carecchio et al 2019 Takezawa et al 2018 30103967 31216378 30682224 27278281 29758257 28668776 29935962 30642806 30103967 30103967 25966631 27068059 27068059 27068059 27068059 28628939 28628939 28668776 31076915 31076915 29761117 Gerald et al 2017 https://doi.org/10.1016/j.spen.2017.08.008 Kelly et al 2019 Koy et al 2018 Danti et al 2017 30682224 30103967 28357411 N N N N Y N N N Y Y N N Y N Y N N Y N N N N N N N N Y Y Y N N N Y N 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61# 62 63 64 65 66 67 68 69 70 71 72* 73* 74 75 76 77 78 79 80 81 NA clonazepam, valproic acid clonidine, clonazepam clonazepam, tetrabenazine, trihexyphenidyl dexmedetomidine, opioids, benzodiazepam, vecuronium, risperidone NA L-DOPA NA clonidine, valproic acid, clonazepam, bethanechol, lorazepam, trihexyphenidyl, dexmedetomidine, propofol, midazolam diazepam, midazolam, clonazepam, phenobarbital, bromazepam, haloperidol, tiapride, eperisone, topiramate NA NA carbamazepine, acetazolamide, oxcarbazepine levodopa, clonazepam, trihexyphenidyl NA haloperidol, midazolam, propofol, phenytoin, baclofen, clonazepam, levodopa, tetrabenazine, clonidine, phenobarbital, lorazepam, curare NA NA risperidone NA N NA N topiramate well-controlled on AED well-controlled on AED carbamazepine, acetazolamide, oxcarbazepine controlled epileptic activity N NA NA N tetrabenazine trihexyphenidyl, clonazepam trihexyphenidyl, clonazepam NA NA midazolam, fentanyl, pimozide, clonazepam, haploperidol, carbamazepine, acetazolamide, diazepam, ketogenic diet levetiracetam,clonazepam, tetrabenazine, respiridone baclofen, phenobarbitone, tetrabenazine midazolam, fentanyl, pimozide levetiracetam, tetrabenazine incomplete response to tetrabenazine NA NA NA NA NA N clonazepam, clonidine, trazodone, midazolam, risperidone, tetrabenazine clonazepam, clobazem, topiramate, levetiracetam, valproic acid, clonidine, baclofen, tetrabenazine, haploperido, diazepam, pentobarbital, propofol, versed, fentanyl, dexmedetomidine NA NA N N NA N tetrabenazine baclofen, clobazem, tetrabenazine, haploperido, diazepam oxcarbazepine, clonazepam, risperidone, diazepam, tetrabenazine diazepam, tetrabenazine L-DOPA levetiracetam N NA haloperidol, baclofen, tetrabenazine, phenobarbital tetrabenazine, phenobarbital Trihexy-phenidyl, nitrazepam, clonazepam, tetrabenazine, baclofen, L-DOPA, lev-etiracetam, phenobarbital Flunitrazepam, baclofen, trihexyphenidyl, tetrabenazine, pimozide NA phenobarbital, clobazam, valproic acid, pyridoxine, levetiracetam, fosphenytoin, valporate, clobazam, zonisamide, rufinamide, high dose prednisone, ACTH, IVIG, ketogenic diet NA NA NA NA NA NA N NA N well-controlled on AED # Patient was presented to our clinic 43 ! N N N NA N N N N Y N N N Y Y N N Y N Y N N Y N N N N N N N N Y Y Y N N N Y N * Patients are siblings Newly reported cases after Feng et al 2018 was publish Figure S1.1 Correlation between seizure frequency and a severe EEG/MRI result Clinical descriptions of EEG and MRI results were classified by one of the authors (C.S.) as normal or not reported ( ), mild (+), or severe (++). The correlation of EEG and MRI findings with either presence or absence of seizures or mutation status are illustrated here and listed in Table S1.1. (A) Over 50% of patients with a seizure disorder (most carrying LOF mutants) exhibit severe EEG. Also, patients carrying the GOF mutation G203R frequently displayed seizure symptoms and severe EEG readings. (B) Patients carrying the GOF mutations G203R and R209C showed occurrence of seizures, and almost half also showed serious abnormalities in MRI. However, less than 50% of the 44 ! Figure S1.1 (cont’d) patients carrying GOF mutation E237K, E246K, or R209H showed seizure activity, while almost 50% of them showed a severe MRI results. All values have been shifted slightly (jitter with SD=5) to avoid overlap. Many patients with singleton LOF mutations fell at 100% Seizures and 100% Severe MRI or EEG. The jitter was added to better demonstrate how many different mutations result in patterns that fall in each region of the graph. 45 ! REFERENCES 46 ! REFERENCES Agler, H. L., Evans, J., Tay, L. H., Anderson, M. J., Colecraft, H. M., & Yue, D. T. (2005). G protein-gated inhibitory module of N-type (ca(v)2.2) ca2+ channels. Neuron, 46(6), 891-904. doi:10.1016/j.neuron.2005.05.011 Ahmad, R., Bourgeois, S., Postnov, A., Schmidt, M. E., Bormans, G., Van Laere, K., & Vandenberghe, W. (2014). PET imaging shows loss of striatal PDE10A in patients with 279-281. doi:10.1212/WNL.0000000000000037 Huntington 82(3), disease. Neurology, Akiyama, H., Fukuda, T., Tojima, T., Nikolaev, V. O., & Kamiguchi, H. (2016). Cyclic Nucleotide Control of Microtubule Dynamics for Axon Guidance. J Neurosci, 36(20), 5636-5649. doi:10.1523/JNEUROSCI.3596-15.2016 Alkufri, F., Shaag, A., Abu-Libdeh, B., & Elpeleg, O. (2016). Deleterious mutation in GPR88 is associated with chorea, speech delay, and learning disabilities. Neurol Genet, 2(3), e64. doi:10.1212/NXG.0000000000000064 Alonso, I., Marques, J. M., Sousa, N., Sequeiros, J., Olsson, I. A., & Silveira, I. (2008). Motor and cognitive deficits in the heterozygous leaner mouse, a Cav2.1 voltage-gated Ca2+ channel mutant. Neurobiol Aging, 29(11), 1733-1743. doi:10.1016/j.neurobiolaging.2007.04.005 Ananth, A. L., Robichaux-Viehoever, A., Kim, Y. M., Hanson-Kahn, A., Cox, R., Enns, G. M., . . . Bernstein, J. A. (2016). Clinical Course of Six Children With GNAO1 Mutations Causing a Severe and Distinctive Movement Disorder. Pediatr Neurol, 59, 81-84. doi:10.1016/j.pediatrneurol.2016.02.018 Antonucci, F., Corradini, I., Fossati, G., Tomasoni, R., Menna, E., & Matteoli, M. (2016). SNAP-25, a Known Presynaptic Protein with Emerging Postsynaptic Functions. Front Synaptic Neurosci, 8, 7. doi:10.3389/fnsyn.2016.00007 Arya, R., Spaeth, C., Gilbert, D. L., Leach, J. L., & Holland, K. D. (2017). GNAO1-associated epileptic encephalopathy and movement disorders: c.607G>A variant represents a probable mutation hotspot with a distinct phenotype. Epileptic Disord, 19(1), 67-75. doi:10.1684/epd.2017.0888 47 ! Baker, K., Gordon, S. L., Grozeva, D., van Kogelenberg, M., Roberts, N. Y., Pike, M., . . . Raymond, F. L. (2015). Identification of a human synaptotagmin-1 mutation that perturbs synaptic vesicle cycling. J Clin Invest, 125(4), 1670-1678. doi:10.1172/JCI79765 Batty, N. J., Fenrich, K. K., & Fouad, K. (2017). The role of cAMP and its downstream targets in neurite growth in the adult nervous system. Neurosci Lett, 652, 56-63. doi:10.1016/j.neulet.2016.12.033 Beaumont, V., & Zucker, R. S. (2000). Enhancement of synaptic transmission by cyclic AMP modulation of presynaptic Ih channels. Nat Neurosci, 3(2), 133-141. doi:10.1038/72072 Bi, Y., Dzierba, C. D., Fink, C., Garcia, Y., Green, M., Han, J., . . . Macor, J. E. (2015). The discovery of potent agonists for GPR88, an orphan GPCR, for the potential treatment of CNS disorders. Bioorg Med Chem Lett, 25(7), 1443-1447. doi:10.1016/j.bmcl.2015.02.038 Blumkin, L., Lerman-Sagie, T., Westenberger, A., Ben-Pazi, H., Zerem, A., Yosovich, K., & Lev, D. (2018). Multiple Causes of Pediatric Early Onset Chorea-Clinical and Genetic Approach. Neuropediatrics, 49(4), 246-255. doi:10.1055/s-0038-1645884 Bollen, E., & Prickaerts, J. (2012). Phosphodiesterases in neurodegenerative disorders. IUBMB Life, 64(12), 965-970. doi:10.1002/iub.1104 Borlikova, G., & Endo, S. (2009). Inducible cAMP early repressor (ICER) and brain functions. Mol Neurobiol, 40(1), 73-86. doi:10.1007/s12035-009-8072-1 Bruun, T. U. J., DesRoches, C. L., Wilson, D., Chau, V., Nakagawa, T., Yamasaki, M., . . . Mercimek-Andrews, S. (2018). Prospective cohort study for identification of underlying genetic causes in neonatal encephalopathy using whole-exome sequencing. Genet Med, 20(5), 486-494. doi:10.1038/gim.2017.129 Carapito, R., Paul, N., Untrau, M., Le Gentil, M., Ott, L., Alsaleh, G., . . . Bahram, S. (2015). A de novo ADCY5 mutation causes early-onset autosomal dominant chorea and dystonia. Mov Disord, 30(3), 423-427. doi:10.1002/mds.26115 Carecchio, M., Invernizzi, F., Gonzalez-Latapi, P., Panteghini, C., Zorzi, G., Romito, L., . . . Nardocci, N. (2019). Frequency and phenotypic spectrum of KMT2B dystonia in childhood: A single-center cohort study. Mov Disord. 48 ! doi:10.1002/mds.27771 Carecchio, M., & Mencacci, N. E. (2017). Emerging Monogenic Complex Hyperkinetic 97. Neurosci 17(12), Rep, Disorders. Neurol doi:10.1007/s11910-017-0806-2 Curr Carecchio, M., Mencacci, N. E., Iodice, A., Pons, R., Panteghini, C., Zorzi, G., . . . Nardocci, N. (2017). ADCY5-related movement disorders: Frequency, disease course and phenotypic variability in a cohort of paediatric patients. Parkinsonism Relat Disord, 41, 37-43. doi:10.1016/j.parkreldis.2017.05.004 Castillo, K., Contreras, G. F., Pupo, A., Torres, Y. P., Neely, A., Gonzalez, C., & Latorre, R. (2015). Molecular mechanism underlying beta1 regulation in voltage- and calcium-activated potassium (BK) channels. Proc Natl Acad Sci U S A, 112(15), 4809-4814. doi:10.1073/pnas.1504378112 Chen, D. H., Meneret, A., Friedman, J. R., Korvatska, O., Gad, A., Bonkowski, E. S., . . . Raskind, W. H. (2015). ADCY5-related dyskinesia: Broader spectrum and correlations. genotype-phenotype 2026-2035. doi:10.1212/WNL.0000000000002058 Neurology, 85(23), Chen, W. J., Lin, Y., Xiong, Z. Q., Wei, W., Ni, W., Tan, G. H., . . . Wu, Z. Y. (2011). that cause kinesigenic dyskinesia. Nat Genet, 43(12), 1252-1255. truncating mutations in PRRT2 identifies Exome sequencing paroxysmal doi:10.1038/ng.1008 Chen, Y. Z., Friedman, J. R., Chen, D. H., Chan, G. C., Bloss, C. S., Hisama, F. M., . . . Torkamani, A. (2014). Gain-of-function ADCY5 mutations in familial dyskinesia with facial myokymia. Ann Neurol, 75(4), 542-549. doi:10.1002/ana.24119 Colecraft, H. M., Brody, D. L., & Yue, D. T. (2001). G-protein inhibition of N- and P/Q-type calcium channels: distinctive elementary mechanisms and their functional impact. J Neurosci, 21(4), 1137-1147. Corradini, I., Verderio, C., Sala, M., Wilson, M. C., & Matteoli, M. (2009). SNAP-25 in 93-99. disorders. Ann N Y Acad Sci, 1152, neuropsychiatric doi:10.1111/j.1749-6632.2008.03995.x Corredor, R. G., Trakhtenberg, E. F., Pita-Thomas, W., Jin, X., Hu, Y., & Goldberg, J. L. (2012). Soluble adenylyl cyclase activity is necessary for retinal ganglion cell 49 ! survival J doi:10.1523/JNEUROSCI.5288-11.2012 growth. axon and Neurosci, 32(22), 7734-7744. D'Angelo, V., Castelli, V., Giorgi, M., Cardarelli, S., Saverioni, I., Palumbo, F., . . . Sancesario, G. (2017). Phosphodiesterase-10A Inverse Changes in Striatopallidal and Striatoentopeduncular Pathways of a Transgenic Mouse Model of DYT1 Dystonia. J Neurosci, 37(8), 2112-2124. doi:10.1523/JNEUROSCI.3207-15.2016 Danti, F. R., Galosi, S., Romani, M., Montomoli, M., Carss, K. J., Raymond, F. L., . . . Guerrini, R. (2017). GNAO1 encephalopathy: Broadening the phenotype and evaluating e143. doi:10.1212/NXG.0000000000000143 outcome. Neurol Genet, treatment 3(2), and de la Fuente-Fernandez, R. (2004). Uncovering the hidden placebo effect in deep-brain stimulation for Parkinson's disease. Parkinsonism Relat Disord, 10(3), 125-127. doi:10.1016/j.parkreldis.2003.10.003 Delgado-Martinez, I., Nehring, R. B., & Sorensen, J. B. (2007). Differential abilities of SNAP-25 homologs to support neuronal function. J Neurosci, 27(35), 9380-9391. doi:10.1523/JNEUROSCI.5092-06.2007 Dhamija, R., Mink, J. W., Shah, B. B., & Goodkin, H. P. (2016). GNAO1-Associated 615-617. Disorder. Mov Disord Pract, 3(6), Movement doi:10.1002/mdc3.12344 Clin Dhamija, R. M., J, W.; Shah, B, B.; Goodkin, H, P. (2016). GNAO1-Associated Movement Disorder. Movement Disorders Clinical Practice, 3(6), 615-617. doi:10.1002/mdc3.12344 Dietel, T. (2016). Genotype and phenotype in GNAO1-Mutation – Case report of an unusual course of a childhood epilepsy. In G. S. T. B. Haack , I. Cordes , R. Trollmann , T. Bast (Ed.), (Vol. 47, pp. P01-17). Neuropediatrics: Thieme. Diggle, C. P., Sukoff Rizzo, S. J., Popiolek, M., Hinttala, R., Schulke, J. P., Kurian, M. A., . . . Brandon, N. J. (2016). Biallelic Mutations in PDE10A Lead to Loss of Striatal PDE10A and a Hyperkinetic Movement Disorder with Onset in Infancy. Am J Hum Genet, 98(4), 735-743. doi:10.1016/j.ajhg.2016.03.015 Dorje, F., Wess, J., Lambrecht, G., Tacke, R., Mutschler, E., & Brann, M. R. (1991). Antagonist binding profiles of five cloned human muscarinic receptor subtypes. J 50 ! Pharmacol Exp Ther, 256(2), 727-733. Dortch-Carnes, J., & Potter, D. E. (2003). Delta-opioid agonist-stimulated inositol phosphate formation in isolated, rabbit iris-ciliary bodies: role of G(i/o) proteins and Gbetagamma-subunits. Exp Eye Res, 77(6), 647-652. Dos Santos, C. O., Masuho, I., da Silva-Junior, F. P., Barbosa, E. R., Silva, S. M., Borges, V., . . . de Carvalho Aguiar, P. (2016). Screening of GNAL variants in Brazilian patients with isolated dystonia reveals a novel mutation with partial loss of function. J Neurol, 263(4), 665-668. doi:10.1007/s00415-016-8026-2 Dufke, C., Sturm, M., Schroeder, C., Moll, S., Ott, T., Riess, O., . . . Grundmann, K. (2014). Screening of mutations in GNAL in sporadic dystonia patients. Mov Disord, 29(9), 1193-1196. doi:10.1002/mds.25794 Eagle, A. L., Gajewski, P. A., & Robison, A. J. (2016). Role of hippocampal activity-induced transcription in memory consolidation. Rev Neurosci, 27(6), 559-573. doi:10.1515/revneuro-2016-0010 Epi, K. C. (2016). De Novo Mutations in SLC1A2 and CACNA1A Are Important Causes of Epileptic Encephalopathies. Am J Hum Genet, 99(2), 287-298. doi:10.1016/j.ajhg.2016.06.003 Epi, K. C., Epilepsy Phenome/Genome, P., Allen, A. S., Berkovic, S. F., Cossette, P., Delanty, N., . . . Winawer, M. R. (2013). De novo mutations in epileptic encephalopathies. Nature, 501(7466), 217-221. doi:10.1038/nature12439 Euro, E.-R. E. S. C., Epilepsy Phenome/Genome, P., & Epi, K. C. (2014). De novo mutations in synaptic transmission genes including DNM1 cause epileptic encephalopathies. 360-370. doi:10.1016/j.ajhg.2014.08.013 Genet, 95(4), Am J Hum Farwell, K. D., Shahmirzadi, L., El-Khechen, D., Powis, Z., Chao, E. C., Tippin Davis, B., . . . Tang, S. (2015). Enhanced utility of family-centered diagnostic exome sequencing with inheritance model-based analysis: results from 500 unselected families with undiagnosed genetic conditions. Genet Med, 17(7), 578-586. doi:10.1038/gim.2014.154 Feng, H., Sjogren, B., Karaj, B., Shaw, V., Gezer, A., & Neubig, R. R. (2017). Movement disorder in GNAO1 encephalopathy associated with gain-of-function mutations. 51 ! Neurology, 89(8), 762-770. doi:10.1212/WNL.0000000000004262 Forbes, E. M., Thompson, A. W., Yuan, J., & Goodhill, G. J. (2012). Calcium and cAMP levels interact to determine attraction versus repulsion in axon guidance. Neuron, 74(3), 490-503. doi:10.1016/j.neuron.2012.02.035 Friedman, J. R., Meneret, A., Chen, D. H., Trouillard, O., Vidailhet, M., Raskind, W. H., & Roze, E. (2016). ADCY5 mutation carriers display pleiotropic paroxysmal day and nighttime dyskinesias. Mov Disord, 31(1), 147-148. doi:10.1002/mds.26494 Fuchs, T., Saunders-Pullman, R., Masuho, I., Luciano, M. S., Raymond, D., Factor, S., . . . Ozelius, L. J. (2013). Mutations in GNAL cause primary torsion dystonia. Nat Genet, 45(1), 88-92. doi:10.1038/ng.2496 Gao, Z., Todorov, B., Barrett, C. F., van Dorp, S., Ferrari, M. D., van den Maagdenberg, A. M., . . . Hoebeek, F. E. (2012). Cerebellar ataxia by enhanced Ca(V)2.1 currents in Cacna1a(S218L) mutant mice. 15533-15546. doi:10.1523/JNEUROSCI.2454-12.2012 is alleviated by Ca2+-dependent K+-channel activators J Neurosci, 32(44), Garza-Lopez, E., Sandoval, A., Gonzalez-Ramirez, R., Gandini, M. A., Van den Maagdenberg, A., De Waard, M., & Felix, R. (2012). Familial hemiplegic migraine type 1 mutations W1684R and V1696I alter G protein-mediated regulation of Ca(V)2.1 voltage-gated calcium channels. Biochim Biophys Acta, 1822(8), 1238-1246. doi:10.1016/j.bbadis.2012.04.008 Gawlinski, P., Posmyk, R., Gambin, T., Sielicka, D., Chorazy, M., Nowakowska, B., . . . Wiszniewski, W. (2016). PEHO Syndrome May Represent Phenotypic Expansion at the Severe End of the Early-Onset Encephalopathies. Pediatr Neurol, 60, 83-87. doi:10.1016/j.pediatrneurol.2016.03.011 Gerald, B., Ramsey, K., Belnap, N., Szelinger, S., Siniard, A. L., Balak, C., . . . Narayanan, V. (2018). Neonatal epileptic encephalopathy caused by de novo GNAO1 mutation misdiagnosed as atypical Rett syndrome: Cautions in interpretation of genomic test results. Semin Pediatr Neurol, 26, 28-32. doi:10.1016/j.spen.2017.08.008 Gill, A., & Hammes, S. R. (2007). G beta gamma signaling reduces intracellular cAMP to promote meiotic progression in mouse oocytes. Steroids, 72(2), 117-123. doi:10.1016/j.steroids.2006.11.006 52 ! Giorgi, M., Melchiorri, G., Nuccetelli, V., D'Angelo, V., Martorana, A., Sorge, R., . . . Sancesario, G. (2011). PDE10A and PDE10A-dependent cAMP catabolism are dysregulated oppositely in striatum and nucleus accumbens after lesion of midbrain dopamine neurons in rat: a key step in parkinsonism physiopathology. Neurobiol Dis, 43(1), 293-303. doi:10.1016/j.nbd.2011.04.006 Goldstein, L. B. (2006). Neurotransmitters and motor activity: effects on functional 451-457. NeuroRx, injury. 3(4), recovery doi:10.1016/j.nurx.2006.07.010 brain after Gorman, K. M., Meyer, E., Grozeva, D., Spinelli, E., McTague, A., Sanchis-Juan, A., . . . Kurian, M. A. (2019). Bi-allelic Loss-of-Function CACNA1B Mutations in Progressive Epilepsy-Dyskinesia. Am J Hum Genet, 104(5), 948-956. doi:10.1016/j.ajhg.2019.03.005 Grimminger, T., Pernhorst, K., Surges, R., Niehusmann, P., Priebe, L., von Lehe, M., . . . Becker, A. J. (2013). Levetiracetam resistance: Synaptic signatures & corresponding promoter SNPs in epileptic hippocampi. Neurobiol Dis, 60, 115-125. doi:10.1016/j.nbd.2013.08.015 Groen, J., van Rootselaar, A. F., van der Salm, S. M., Bloem, B. R., & Tijssen, M. (2011). A new familial syndrome with dystonia and lower limb action myoclonus. Mov Disord, 26(5), 896-900. doi:10.1002/mds.23557 Groen, J. L., Andrade, A., Ritz, K., Jalalzadeh, H., Haagmans, M., Bradley, T. E., . . . Tijssen, M. A. (2015). CACNA1B mutation is linked to unique myoclonus-dystonia syndrome. Hum Mol Genet, 24(4), 987-993. doi:10.1093/hmg/ddu513 Guan, J. S., Su, S. C., Gao, J., Joseph, N., Xie, Z., Zhou, Y., . . . Tsai, L. H. (2011). Cdk5 is required for memory function and hippocampal plasticity via the cAMP signaling pathway. PLoS One, 6(9), e25735. doi:10.1371/journal.pone.0025735 Guida, S., Trettel, F., Pagnutti, S., Mantuano, E., Tottene, A., Veneziano, L., . . . Pietrobon, D. (2001). Complete loss of P/Q calcium channel activity caused by a CACNA1A missense mutation carried by patients with episodic ataxia type 2. Am J Hum Genet, 68(3), 759-764. Helbig, K. L., Farwell Hagman, K. D., Shinde, D. N., Mroske, C., Powis, Z., Li, S., . . . Helbig, I. (2016). Diagnostic exome sequencing provides a molecular diagnosis for a significant proportion of patients with epilepsy. Genet Med, 18(9), 898-905. 53 ! doi:10.1038/gim.2015.186 Heron, S. E., Grinton, B. E., Kivity, S., Afawi, Z., Zuberi, S. M., Hughes, J. N., . . . Dibbens, L. M. (2012). PRRT2 mutations cause benign familial infantile epilepsy and infantile convulsions with choreoathetosis syndrome. Am J Hum Genet, 90(1), 152-160. doi:10.1016/j.ajhg.2011.12.003 Honey, C. M., Malhotra, A. K., Tarailo-Graovac, M., van Karnebeek, C. D. M., Horvath, G., & Sulistyanto, A. (2018). GNAO1 Mutation-Induced Pediatric Dystonic Storm Rescue With Pallidal Deep Brain Stimulation. J Child Neurol, 33(6), 413-416. doi:10.1177/0883073818756134 Inda, C., Bonfiglio, J. J., Dos Santos Claro, P. A., Senin, S. A., Armando, N. G., Deussing, J. M., & Silberstein, S. (2017). cAMP-dependent cell differentiation triggered by activated CRHR1 in hippocampal neuronal cells. Sci Rep, 7(1), 1944. doi:10.1038/s41598-017-02021-7 Iwamoto, T., Okumura, S., Iwatsubo, K., Kawabe, J., Ohtsu, K., Sakai, I., . . . Ishikawa, Y. (2003). Motor dysfunction in type 5 adenylyl cyclase-null mice. J Biol Chem, 278(19), 16936-16940. doi:10.1074/jbc.C300075200 Jen, J. C., Yue, Q., Karrim, J., Nelson, S. F., & Baloh, R. W. (1998). Spinocerebellar ataxia type 6 with positional vertigo and acetazolamide responsive episodic ataxia. J Neurol Neurosurg Psychiatry, 65(4), 565-568. Jeng, C. J., Chen, Y. T., Chen, Y. W., & Tang, C. Y. (2006). Dominant-negative effects of human P/Q-type Ca2+ channel mutations associated with episodic ataxia type 2. Am J Physiol Cell Physiol, 290(4), C1209-1220. doi:10.1152/ajpcell.00247.2005 Jeon, C. Y., Moon, M. Y., Kim, J. H., Kim, H. J., Kim, J. G., Li, Y., . . . Park, J. B. (2012). Control of neurite outgrowth by RhoA inactivation. J Neurochem, 120(5), 684-698. doi:10.1111/j.1471-4159.2011.07564.x Jiang, M., & Bajpayee, N. S. (2009). Molecular mechanisms of go signaling. Neurosignals, 17(1), 23-41. doi:10.1159/000186688 Jiang, M., Gold, M. S., Boulay, G., Spicher, K., Peyton, M., Brabet, P., . . . Birnbaumer, L. (1998). Multiple neurological abnormalities in mice deficient in the G protein Go. Proc Natl Acad Sci U S A, 95(6), 3269-3274. doi:10.1073/pnas.95.6.3269 54 ! Jin, C., Decker, A. M., Huang, X. P., Gilmour, B. P., Blough, B. E., Roth, B. L., . . . Zhang, X. P. (2014). Synthesis, pharmacological characterization, and structure-activity relationship studies of small molecular agonists for the orphan GPR88 receptor. ACS Chem Neurosci, 5(7), 576-587. doi:10.1021/cn500082p Jouvenceau, A., Eunson, L. H., Spauschus, A., Ramesh, V., Zuberi, S. M., Kullmann, D. M., & Hanna, M. G. (2001). Human epilepsy associated with dysfunction of the brain 801-807. doi:10.1016/S0140-6736(01)05971-2 358(9284), P/Q-type channel. calcium Lancet, Kandel, E. R. (2012). The molecular biology of memory: cAMP, PKA, CRE, CREB-1, CREB-2, and CPEB. Mol Brain, 5, 14. doi:10.1186/1756-6606-5-14 Kehrl, J. M., Sahaya, K., Dalton, H. M., Charbeneau, R. A., Kohut, K. T., Gilbert, K., . . . Neubig, R. R. (2014). Gain-of-function mutation in Gnao1: a murine model of epileptiform encephalopathy (EIEE17)? Mamm Genome, 25(5-6), 202-210. doi:10.1007/s00335-014-9509-z Kelly, M., Park, M., Mihalek, I., Rochtus, A., Gramm, M., Perez-Palma, E., . . . Poduri, A. (2019). Spectrum of neurodevelopmental disease associated with the GNAO1 guanosine 406-418. doi:10.1111/epi.14653 triphosphate-binding Epilepsia, region. 60(3), Koy, A., Cirak, S., Gonzalez, V., Becker, K., Roujeau, T., Milesi, C., . . . Cif, L. (2018). Deep brain stimulation is effective in pediatric patients with GNAO1 associated severe hyperkinesia. J Neurol Sci, 391, 31-39. doi:10.1016/j.jns.2018.05.018 Kulkarni, N., Tang, S., Bhardwaj, R., Bernes, S., & Grebe, T. A. (2016). Progressive Movement Disorder in Brothers Carrying a GNAO1 Mutation Responsive to Deep Brain 211-214. doi:10.1177/0883073815587945 Stimulation. Neurol, 31(2), J Child Kumar, K. R., Lohmann, K., Masuho, I., Miyamoto, R., Ferbert, A., Lohnau, T., . . . Schmidt, A. (2014). Mutations in GNAL: a novel cause of craniocervical dystonia. JAMA Neurol, 71(4), 490-494. doi:10.1001/jamaneurol.2013.4677 Law, C. Y., Chang, S. T., Cho, S. Y., Yau, E. K., Ng, G. S., Fong, N. C., & Lam, C. W. (2015). Clinical whole-exome sequencing reveals a novel missense pathogenic variant of GNAO1 in a patient with infantile-onset epilepsy. Clin Chim Acta, 451(Pt B), 292-296. doi:10.1016/j.cca.2015.10.011 55 ! Law, C. Y., Yeung, W. L., Cheung, Y. F., Chan, H. F., Fung, E., Hui, J., . . . Lam, C. W. (2016). A common PRRT2 mutation in familial paroxysmal kinesigenic dyskinesia in Hong Kong: a case series of 16 patients. Hong Kong Med J, 22(6), 619-622. doi:10.12809/hkmj154579 Lee, H. Y., Huang, Y., Bruneau, N., Roll, P., Roberson, E. D., Hermann, M., . . . Ptacek, L. J. (2012). Mutations in the gene PRRT2 cause paroxysmal kinesigenic dyskinesia 2-12. doi:10.1016/j.celrep.2011.11.001 convulsions. infantile with Cell Rep, 1(1), Lee, U. S., & Cui, J. (2010). BK channel activation: structural and functional insights. Trends Neurosci, 33(9), 415-423. doi:10.1016/j.tins.2010.06.004 Lepski, G., Jannes, C. E., Nikkhah, G., & Bischofberger, J. (2013). cAMP promotes the differentiation of neural progenitor cells in vitro via modulation of voltage-gated calcium channels. Front Cell Neurosci, 7, 155. doi:10.3389/fncel.2013.00155 Lessmann, V., & Heumann, R. (1997). Cyclic AMP endogenously enhances synaptic strength of developing glutamatergic synapses in serum-free microcultures of rat hippocampal neurons. Brain Res, 763(1), 111-122. Leung, C. C. Y., & Wong, Y. H. (2017). Role of G Protein-Coupled Receptors in the Regulation of Structural Plasticity and Cognitive Function. Molecules, 22(7). doi:10.3390/molecules22071239 Leveque, C., el Far, O., Martin-Moutot, N., Sato, K., Kato, R., Takahashi, M., & Seagar, M. J. (1994). Purification of the N-type calcium channel associated with syntaxin and synaptotagmin. A complex implicated in synaptic vesicle exocytosis. J Biol Chem, 269(9), 6306-6312. Lohmann, K., Masuho, I., Patil, D. N., Baumann, H., Hebert, E., Steinrucke, S., . . . Martemyanov, K. A. (2017). Novel GNB1 mutations disrupt assembly and function of G protein heterotrimers and cause global developmental delay in humans. Hum Mol Genet, 26(6), 1078-1086. doi:10.1093/hmg/ddx018 Luscher, C., & Slesinger, P. A. (2010). Emerging roles for G protein-gated inwardly rectifying potassium (GIRK) channels in health and disease. Nat Rev Neurosci, 11(5), 301-315. doi:10.1038/nrn2834 Malaquias, M. J., Fineza, I., Loureiro, L., Cardoso, L., Alonso, I., & Magalhaes, M. (2019). 56 ! GNAO1 mutation presenting as dyskinetic cerebral palsy. Neurol Sci. doi:10.1007/s10072-019-03964-7 Marce-Grau, A., Dalton, J., Lopez-Pison, J., Garcia-Jimenez, M. C., Monge-Galindo, L., Cuenca-Leon, E., . . . Macaya, A. (2016). GNAO1 encephalopathy: further delineation of a severe neurodevelopmental syndrome affecting females. Orphanet J Rare Dis, 11, 38. doi:10.1186/s13023-016-0416-0 Massart, R., Guilloux, J. P., Mignon, V., Sokoloff, P., & Diaz, J. (2009). Striatal GPR88 expression is confined to the whole projection neuron population and is regulated by dopaminergic and glutamatergic afferents. Eur J Neurosci, 30(3), 397-414. doi:10.1111/j.1460-9568.2009.06842.x Masuho, I., Fang, M., Geng, C., Zhang, J., Jiang, H., Ozgul, R. K., . . . Kruer, M. C. (2016). Homozygous GNAL mutation associated with familial childhood-onset generalized e78. doi:10.1212/NXG.0000000000000078 dystonia. 2(3), Neurol Genet, Mayfield, J., Blednov, Y. A., & Harris, R. A. (2015). Behavioral and Genetic Evidence for GIRK Channels in the CNS: Role in Physiology, Pathophysiology, and Drug Addiction. Int Rev Neurobiol, 123, 279-313. doi:10.1016/bs.irn.2015.05.016 McDavid, S., & Currie, K. P. (2006). G-proteins modulate cumulative inactivation of (Cav2.2) calcium channels. J Neurosci, 26(51), 13373-13383. N-type doi:10.1523/JNEUROSCI.3332-06.2006 McIntyre, C. C., & Anderson, R. W. (2016). Deep brain stimulation mechanisms: the control of network activity via neurochemistry modulation. J Neurochem, 139 Suppl 1, 338-345. doi:10.1111/jnc.13649 Meijer, I. A., Miravite, J., Kopell, B. H., & Lubarr, N. (2017). Deep Brain Stimulation in an Additional Patient With ADCY5-Related Movement Disorder. J Child Neurol, 32(4), 438-439. doi:10.1177/0883073816681353 Meirsman, A. C., Le Merrer, J., Pellissier, L. P., Diaz, J., Clesse, D., Kieffer, B. L., & Becker, J. A. (2016). Mice Lacking GPR88 Show Motor Deficit, Improved Spatial Learning, and Low Anxiety Reversed by Delta Opioid Antagonist. Biol Psychiatry, 79(11), 917-927. doi:10.1016/j.biopsych.2015.05.020 Mencacci, N. E., & Carecchio, M. (2016). Recent advances in genetics of chorea. Curr 57 ! Opin Neurol, 29(4), 486-495. doi:10.1097/WCO.0000000000000352 Mencacci, N. E., Erro, R., Wiethoff, S., Hersheson, J., Ryten, M., Balint, B., . . . Bhatia, K. P. (2015). ADCY5 mutations are another cause of benign hereditary chorea. Neurology, 85(1), 80-88. doi:10.1212/WNL.0000000000001720 Mencacci, N. E., Kamsteeg, E. J., Nakashima, K., R'Bibo, L., Lynch, D. S., Balint, B., . . . Bhatia, K. P. (2016). De Novo Mutations in PDE10A Cause Childhood-Onset Chorea with Bilateral Striatal Lesions. Am J Hum Genet, 98(4), 763-771. doi:10.1016/j.ajhg.2016.02.015 Mencacci, N. E., R'Bibo, L., Bandres-Ciga, S., Carecchio, M., Zorzi, G., Nardocci, N., . . . Wood, N. W. (2015). The CACNA1B R1389H variant is not associated with myoclonus-dystonia in a large European multicentric cohort. Hum Mol Genet, 24(18), 5326-5329. doi:10.1093/hmg/ddv255 Menke, L. A., Engelen, M., Alders, M., Odekerken, V. J., Baas, F., & Cobben, J. M. (2016). Recurrent GNAO1 Mutations Associated With Developmental Delay and a Movement 1598-1601. doi:10.1177/0883073816666474 Disorder. 31(14), J Child Neurol, Mercado, R., Constantoyannis, C., Mandat, T., Kumar, A., Schulzer, M., Stoessl, A. J., & Honey, C. R. (2006). Expectation and the placebo effect in Parkinson's disease patients with subthalamic nucleus deep brain stimulation. Mov Disord, 21(9), 1457-1461. doi:10.1002/mds.20935 Mercimek-Mahmutoglu, S., Sidky, S., Hyland, K., Patel, J., Donner, E. J., Logan, W., . . . Kyriakopoulou, L. (2015). Prevalence of inherited neurotransmitter disorders in patients with movement disorders and epilepsy: a retrospective cohort study. Orphanet J Rare Dis, 10, 12. doi:10.1186/s13023-015-0234-9 Miclea, D., Peca, L., Cuzmici, Z., & Pop, I. V. (2015). Genetic testing in patients with global developmental delay / intellectual disabilities. A review. Clujul Med, 88(3), 288-292. doi:10.15386/cjmed-461 Mizushima, K., Miyamoto, Y., Tsukahara, F., Hirai, M., Sakaki, Y., & Ito, T. (2000). A novel G-protein-coupled receptor gene expressed in striatum. Genomics, 69(3), 314-321. doi:10.1006/geno.2000.6340 Mullner, C., Broos, L. A., van den Maagdenberg, A. M., & Striessnig, J. (2004). Familial 58 ! hemiplegic migraine type 1 mutations K1336E, W1684R, and V1696I alter Cav2.1 Ca2+ channel gating: evidence for beta-subunit isoform-specific effects. J Biol Chem, 279(50), 51844-51850. doi:10.1074/jbc.M408756200 Nakamura, K., Kodera, H., Akita, T., Shiina, M., Kato, M., Hoshino, H., . . . Saitsu, H. (2013). De Novo mutations in GNAO1, encoding a Galphao subunit of heterotrimeric G proteins, cause epileptic encephalopathy. Am J Hum Genet, 93(3), 496-505. doi:10.1016/j.ajhg.2013.07.014 Neher, E. (2006). A comparison between exocytic control mechanisms in adrenal chromaffin cells and a glutamatergic synapse. Pflugers Arch, 453(3), 261-268. doi:10.1007/s00424-006-0143-9 Nestler, E. J., Alreja, M., & Aghajanian, G. K. (1999). Molecular control of locus coeruleus neurotransmission. Biol Psychiatry, 46(9), 1131-1139. Ohno, Y., & Tokudome, K. (2017). Therapeutic Role of Synaptic Vesicle Glycoprotein 2A (SV2A) in Modulating Epileptogenesis. CNS Neurol Disord Drug Targets, 16(4), 463-471. doi:10.2174/1871527316666170404115027 Okumura, A., Maruyama, K., Shibata, M., Kurahashi, H., Ishii, A., Numoto, S., . . . Kojima, S. (2018). A patient with a GNAO1 mutation with decreased spontaneous movements, hypotonia, and dystonic features. Brain Dev, 40(10), 926-930. doi:10.1016/j.braindev.2018.06.005 Paddock, B. E., Wang, Z., Biela, L. M., Chen, K., Getzy, M. D., Striegel, A., . . . Reist, N. E. (2011). Membrane penetration by synaptotagmin is required for coupling calcium binding to vesicle fusion in vivo. J Neurosci, 31(6), 2248-2257. doi:10.1523/JNEUROSCI.3153-09.2011 Padovan-Neto, F. E., & West, A. R. (2017). Regulation of Striatal Neuron Activity by Cyclic Nucleotide Signaling and Phosphodiesterase Inhibition: Implications for the Treatment 257-283. doi:10.1007/978-3-319-58811-7_10 of Parkinson's Disease. Adv Neurobiol, 17, Park, H. Y., Kang, Y. M., Kang, Y., Park, T. S., Ryu, Y. K., Hwang, J. H., . . . Kim, K. S. (2014). Inhibition of adenylyl cyclase type 5 prevents L-DOPA-induced dyskinesia in an animal model of Parkinson's disease. J Neurosci, 34(35), 11744-11753. doi:10.1523/JNEUROSCI.0864-14.2014 59 ! Patel, T. B. (2004). Single transmembrane spanning heterotrimeric g protein-coupled receptors and their signaling cascades. Pharmacol Rev, 56(3), 371-385. doi:10.1124/pr.56.3.4 Pelosi, A., Menardy, F., Popa, D., Girault, J. A., & Herve, D. (2017). Heterozygous Gnal Mice Are a Novel Animal Model with Which to Study Dystonia Pathophysiology. J Neurosci, 37(26), 6253-6267. doi:10.1523/JNEUROSCI.1529-16.2017 Petrovski, S., Kury, S., Myers, C. T., Anyane-Yeboa, K., Cogne, B., Bialer, M., . . . Goldstein, D. B. (2016). Germline De Novo Mutations in GNB1 Cause Severe Neurodevelopmental Disability, Hypotonia, and Seizures. Am J Hum Genet, 98(5), 1001-1010. doi:10.1016/j.ajhg.2016.03.011 Pierre, S., Eschenhagen, T., Geisslinger, G., & Scholich, K. (2009). Capturing adenylyl cyclases as potential drug targets. Nat Rev Drug Discov, 8(4), 321-335. doi:10.1038/nrd2827 Putzel, G. G., Fuchs, T., Battistella, G., Rubien-Thomas, E., Frucht, S. J., Blitzer, A., . . . Simonyan, K. (2016). GNAL mutation in isolated laryngeal dystonia. Mov Disord, 31(5), 750-755. doi:10.1002/mds.26502 Quintana, A., Sanz, E., Wang, W., Storey, G. P., Guler, A. D., Wanat, M. J., . . . Palmiter, R. D. (2012). Lack of GPR88 enhances medium spiny neuron activity and alters motor- and cue-dependent behaviors. Nat Neurosci, 15(11), 1547-1555. doi:10.1038/nn.3239 Rickman, C., & Davletov, B. (2003). Mechanism of calcium-independent synaptotagmin 5501-5504. target SNAREs. J Biol Chem, binding doi:10.1074/jbc.C200692200 to 278(8), Rohena, L., Neidich, J., Truitt Cho, M., Gonzalez, K. D., Tang, S., Devinsky, O., & Chung, W. K. (2013). Mutation in SNAP25 as a novel genetic cause of epilepsy and intellectual disability. Rare Dis, 1, e26314. doi:10.4161/rdis.26314 Russwurm, C., Koesling, D., & Russwurm, M. (2015). Phosphodiesterase 10A Is Tethered to a Synaptic Signaling Complex in Striatum. J Biol Chem, 290(19), 11936-11947. doi:10.1074/jbc.M114.595769 Saitsu, H., Fukai, R., Ben-Zeev, B., Sakai, Y., Mimaki, M., Okamoto, N., . . . Matsumoto, N. (2016). Phenotypic spectrum of GNAO1 variants: epileptic encephalopathy to 60 ! involuntary movements with severe developmental delay. Eur J Hum Genet, 24(1), 129-134. doi:10.1038/ejhg.2015.92 Sakamoto, S., Monden, Y., Fukai, R., Miyake, N., Saito, H., Miyauchi, A., . . . Yamagata, T. (2017). A case of severe movement disorder with GNAO1 mutation responsive to topiramate. Brain Dev, 39(5), 439-443. doi:10.1016/j.braindev.2016.11.009 Salvi, J., Bertaso, F., Mausset-Bonnefont, A. L., Metz, A., Lemmers, C., Ango, F., . . . Mezghrani, A. (2014). RNAi silencing of P/Q-type calcium channels in Purkinje neurons of adult mouse leads to episodic ataxia type 2. Neurobiol Dis, 68, 47-56. doi:10.1016/j.nbd.2014.04.005 Schirinzi, T., Garone, G., Travaglini, L., Vasco, G., Galosi, S., Rios, L., . . . Leuzzi, V. (2019). Phenomenology and clinical course of movement disorder in GNAO1 variants: Results from an analytical review. Parkinsonism Relat Disord, 61, 19-25. doi:10.1016/j.parkreldis.2018.11.019 Schmidt, M., Dekker, F. J., & Maarsingh, H. (2013). Exchange protein directly activated by cAMP (epac): a multidomain cAMP mediator in the regulation of diverse biological functions. Pharmacol Rev, 65(2), 670-709. doi:10.1124/pr.110.003707 Schorling, D. C., Dietel, T., Evers, C., Hinderhofer, K., Korinthenberg, R., Ezzo, D., . . . Kirschner, J. (2017). Expanding Phenotype of De Novo Mutations in GNAO1: Four New Cases and Review of Literature. Neuropediatrics, 48(5), 371-377. doi:10.1055/s-0037-1603977 Shanley, L. J., O'Malley, D., Irving, A. J., Ashford, M. L., & Harvey, J. (2002). Leptin inhibits epileptiform-like activity rat hippocampal neurones via PI 3-kinase-driven activation of BK channels. J Physiol, 545(3), 933-944. doi:10.1113/jphysiol.2002.029488 in Sharma, S., Hansen, J. T., & Notter, M. F. (1990). Effects of NGF and dibutyryl cAMP on neuronal differentiation of embryonal carcinoma cells. Int J Dev Neurosci, 8(1), 33-45. Shaw, C., Hisama, F., Friedman, J., & Bird, T. D. (1993). ADCY5-Related Dyskinesia. In M. P. Adam, H. H. Ardinger, R. A. Pagon, S. E. Wallace, L. J. H. Bean, K. Stephens, & A. Amemiya (Eds.), GeneReviews((R)). Seattle (WA). Signorini, S., Liao, Y. J., Duncan, S. A., Jan, L. Y., & Stoffel, M. (1997). Normal 61 ! cerebellar development but susceptibility to seizures in mice lacking G protein-coupled, inwardly rectifying K+ channel GIRK2. Proc Natl Acad Sci U S A, 94(3), 923-927. doi:10.1073/pnas.94.3.923 Simms, B. A., & Zamponi, G. W. (2014). Neuronal voltage-gated calcium channels: 24-45. dysfunction. Neuron, 82(1), and structure, doi:10.1016/j.neuron.2014.03.016 function, Sintas, C., Carreno, O., Fernandez-Castillo, N., Corominas, R., Vila-Pueyo, M., Toma, C., . . . Cormand, B. (2017). Mutation Spectrum in the CACNA1A Gene in 49 Patients 2514. doi:10.1038/s41598-017-02554-x Episodic Ataxia. Rep, 7(1), with Sci Siuciak, J. A., McCarthy, S. A., Chapin, D. S., Martin, A. N., Harms, J. F., & Schmidt, C. J. (2008). the phosphodiesterase-10A (PDE10A) enzyme on a C57/Bl6N congenic background. Neuropharmacology, 54(2), 417-427. doi:10.1016/j.neuropharm.2007.10.009 characterization Behavioral of mice deficient in Steinrucke, S., Lohmann, K., Domingo, A., Rolfs, A., Baumer, T., Spiegler, J., . . . Munchau, A. (2016). Novel GNB1 missense mutation in a patient with generalized dystonia, hypotonia, and intellectual disability. Neurol Genet, 2(5), e106. doi:10.1212/NXG.0000000000000106 Sudhof, T. C. (2012). Calcium control of neurotransmitter release. Cold Spring Harb Perspect Biol, 4(1), a011353. doi:10.1101/cshperspect.a011353 Sutton, K. G., McRory, J. E., Guthrie, H., Murphy, T. H., & Snutch, T. P. (1999). P/Q-type calcium channels mediate the activity-dependent feedback of syntaxin-1A. Nature, 401(6755), 800-804. doi:10.1038/44586 Szczepanik, E., Terczynska, I., Kruk, M., Lipiec, A., Dudko, E., Tryfon, J., . . . Hoffman-Zacharska, D. (2015). Glucose transporter type 1 deficiency due to SLC2A1 gene mutations--a rare but treatable cause of metabolic epilepsy and extrapyramidal movement disorder; own experience and literature review. Dev Period Med, 19(4), 454-463. Tabarki, B., AlMajhad, N., AlHashem, A., Shaheen, R., & Alkuraya, F. S. (2016). Homozygous KCNMA1 mutation as a cause of cerebellar atrophy, developmental delay 1295-1298. doi:10.1007/s00439-016-1726-y seizures. 135(11), Genet, Hum and 62 ! Takezawa, Y., Kikuchi, A., Haginoya, K., Niihori, T., Numata-Uematsu, Y., Inui, T., . . . Kure, S. (2018). Genomic analysis identifies masqueraders of full-term cerebral palsy. Ann Clin Transl Neurol, 5(5), 538-551. doi:10.1002/acn3.551 Talvik, I., Moller, R. S., Vaher, M., Vaher, U., Larsen, L. H., Dahl, H. A., . . . Talvik, T. (2015). Clinical Phenotype of De Novo GNAO1 Mutation: Case Report and Review of Literature. Child Neurol Open, 2(2), 2329048X15583717. doi:10.1177/2329048X15583717 Tong, J., Liu, X., Vickstrom, C., Li, Y., Yu, L., Lu, Y., . . . Liu, Q. S. (2017). The Epac-Phospholipase Cepsilon Pathway Regulates Endocannabinoid Signaling and Cocaine-Induced Disinhibition of Ventral Tegmental Area Dopamine Neurons. J Neurosci, 37(11), 3030-3044. doi:10.1523/JNEUROSCI.2810-16.2017 Torrecilla, M., Marker, C. L., Cintora, S. C., Stoffel, M., Williams, J. T., & Wickman, K. (2002). G-protein-gated potassium channels containing Kir3.2 and Kir3.3 subunits mediate the acute inhibitory effects of opioids on locus ceruleus neurons. J Neurosci, 22(11), 4328-4334. doi:20026414 Tucker, W. C., Weber, T., & Chapman, E. R. (2004). Reconstitution of Ca2+-regulated membrane fusion by synaptotagmin and SNAREs. Science, 304(5669), 435-438. doi:10.1126/science.1097196 Ueda, K., Serajee, F., & Huq, A. (2016). Exome Sequencing Identifies Dual Mutations in Calcium Signaling Genes GNAO1 and ATP2B3 in a Patient with Early Infantile Epileptic Encephalopathy (EIEE) (P5.146). Neurology, 86. Waak, M., Mohammad, S. S., Coman, D., Sinclair, K., Copeland, L., Silburn, P., . . . Malone, S. (2017). GNAO1-related movement disorder with life-threatening exacerbations: movement phenomenology and response to DBS. J Neurol Neurosurg Psychiatry. doi:10.1136/jnnp-2017-315653 Waak, M., Mohammad, S. S., Coman, D., Sinclair, K., Copeland, L., Silburn, P., . . . Malone, S. (2018). GNAO1-related movement disorder with life-threatening exacerbations: movement phenomenology and response to DBS. J Neurol Neurosurg Psychiatry, 89(2), 221-222. doi:10.1136/jnnp-2017-315653 Wan, J., Mamsa, H., Johnston, J. L., Spriggs, E. L., Singer, H. S., Zee, D. S., . . . Investigators, C. (2011). Large Genomic Deletions in CACNA1A Cause Episodic Ataxia Type 2. Front Neurol, 2, 51. doi:10.3389/fneur.2011.00051 63 ! Wang, J. L., Cao, L., Li, X. H., Hu, Z. M., Li, J. D., Zhang, J. G., . . . Tang, B. S. (2011). Identification of PRRT2 as the causative gene of paroxysmal kinesigenic dyskinesias. Brain, 134(Pt 12), 3493-3501. doi:10.1093/brain/awr289 Westenberger, A., Max, C., Bruggemann, N., Domingo, A., Grutz, K., Pawlack, H., . . . Klein, C. (2017). Alternating Hemiplegia of Childhood as a New Presentation of Adenylate Cyclase 5-Mutation-Associated Disease: A Report of Two Cases. J Pediatr, 181, 306-308 e301. doi:10.1016/j.jpeds.2016.10.079 Weston, M. C. (2017). The Role of PRRT2 in Synaptic Transmission May Not Be So Benign. Epilepsy Curr, 17(3), 165-166. doi:10.5698/1535-7511.17.3.165 Worley, P. F., Baraban, J. M., Van Dop, C., Neer, E. J., & Snyder, S. H. (1986). Go, a guanine nucleotide-binding protein: immunohistochemical localization in rat brain resembles distribution of second messenger systems. Proc Natl Acad Sci U S A, 83(12), 4561-4565. doi:10.1073/pnas.83.12.4561 Xie, Z., Adamowicz, W. O., Eldred, W. D., Jakowski, A. B., Kleiman, R. J., Morton, D. G., . . . Menniti, F. S. (2006). Cellular and subcellular localization of PDE10A, a striatum-enriched 597-607. doi:10.1016/j.neuroscience.2005.12.042 phosphodiesterase. Neuroscience, 139(2), Xiong, J., Peng, J., Duan, H. L., Chen, C., Wang, X. L., Chen, S. M., & Yin, F. (2018). [Recurrent convulsion and pulmonary infection complicated by psychomotor retardation in an infant]. Zhongguo Dang Dai Er Ke Za Zhi, 20(2), 154-157. Yesil, G., Aralasmak, A., Akyuz, E., Icagasioglu, D., Uygur Sahin, T., & Bayram, Y. (2018). Expanding the Phenotype of Homozygous KCNMA1 Mutations; Dyskinesia, Epilepsy, Intellectual Disability, Cerebellar and Corticospinal Tract Atrophy. Balkan Med J, 35(4), 336-339. doi:10.4274/balkanmedj.2017.0986 Yilmaz, S., Turhan, T., Ceylaner, S., Gokben, S., Tekgul, H., & Serdaroglu, G. (2016). Excellent response to deep brain stimulation in a young girl with GNAO1-related progressive 1567-1568. doi:10.1007/s00381-016-3139-6 choreoathetosis. Childs Nerv Syst, 32(9), Yoda, A., Adelmant, G., Tamburini, J., Chapuy, B., Shindoh, N., Yoda, Y., . . . Lane, A. A. (2015). Mutations in G protein beta subunits promote transformation and kinase inhibitor resistance. Nat Med, 21(1), 71-75. doi:10.1038/nm.3751 64 ! Yu, J., Wang, S., Wu, C., & Yi, S. (2017). Deep Sequencing Reveals the Significant Involvement of cAMP-Related Signaling Pathways Following Sciatic Nerve Crush. Neurochem Res, 42(12), 3603-3611. doi:10.1007/s11064-017-2409-3 Zamponi, G. W., & Currie, K. P. (2013). Regulation of Ca(V)2 calcium channels by G receptors. Biochim Biophys Acta, 1828(7), 1629-1643. protein coupled doi:10.1016/j.bbamem.2012.10.004 Zhang, Z. B., Tian, M. Q., Gao, K., Jiang, Y. W., & Wu, Y. (2015). De novo KCNMA1 mutations in children with early-onset paroxysmal dyskinesia and developmental delay. Mov Disord, 30(9), 1290-1292. doi:10.1002/mds.26216 Zhu, X., Petrovski, S., Xie, P., Ruzzo, E. K., Lu, Y. F., McSweeney, K. M., . . . Goldstein, D. B. (2015). Whole-exome sequencing in undiagnosed genetic diseases: interpreting 119 trios. Genet Med, 17(10), 774-781. doi:10.1038/gim.2014.191 Ztaou, S., Maurice, N., Camon, J., Guiraudie-Capraz, G., Kerkerian-Le Goff, L., Beurrier, C., . . . Amalric, M. (2016). Involvement of Striatal Cholinergic Interneurons and M1 and M4 Muscarinic Receptors in Motor Symptoms of Parkinson's Disease. J Neurosci, 36(35), 9161-9172. doi:10.1523/JNEUROSCI.0873-16.2016 Zurawski, Z., Page, B., Chicka, M. C., Brindley, R. L., Wells, C. A., Preininger, A. M., . . . Hamm, H. E. (2017). Gbetagamma directly modulates vesicle fusion by competing with synaptotagmin for binding to neuronal SNARE proteins embedded in 12165-12177. doi:10.1074/jbc.M116.773523 membranes. 292(29), Chem, Biol J Zurawski, Z., Rodriguez, S., Hyde, K., Alford, S., & Hamm, H. E. (2016). Gbetagamma Binds to the Extreme C Terminus of SNAP25 to Mediate the Action of Gi/o-Coupled G Protein-Coupled Receptors. Mol Pharmacol, 89(1), 75-83. doi:10.1124/mol.115.101600 Zwart, R., Reed, H., Clarke, S., & Sher, E. (2016). A novel muscarinic receptor-independent mechanism of KCNQ2/3 potassium channel blockade by Oxotremorine-M. 221-228. doi:10.1016/j.ejphar.2016.08.037 Pharmacol, 791, Eur J 65 ! CHAPTER 2: MOVEMENT DISORDER IN GNAO1 ENCEPHALOPATHY ASSOCIATED WITH GAIN-OF-FUNCTION MUTATIONS Modified from Feng, H., Sjögren, B., Karaj, B., Shaw, V., Gezer, A., & Neubig, R. R. (2017). Movement disorder in GNAO1 encephalopathy associated with gain-of-function mutations. Neurology, 89(8), 762-770. DOI: https://doi.org/10.1212/WNL.0000000000004262 With permission from the American Academy of Neurology. All rights reserved. Gezer, A. performed plasmid mutagenesis for the first set of GNAO1 mutations. Wellhausen, N. did mutagenesis and cAMP assay in Figure S2.8. Karaj, B. prepared six of the GNAO1 mutant plasmids used in Figure 2.2. Vincent Shaw made Figure S2.5. 66! 2.1 Abstract Objective: To define molecular mechanisms underlying the clinical spectrum of epilepsy and movement disorder in individuals with de novo mutations in the GNAO1 gene. Methods: We identified all GNAO1 mutations reported in individuals with epilepsy (EIEE17) or movement disorders through April 2016; 15 de novo mutant alleles from 25 individuals were introduced into the Gαo subunit by site-directed mutagenesis in a mammalian expression plasmid. We assessed protein expression and function in vitro in HEK-293T cells by western blot and determined functional Gαo-dependent cyclic AMP inhibition with a co-expressed α2A adrenergic receptor. Results: Of the 15 clinical GNAO1 mutations studied, 9 show reduced expression and loss of function (LOF, <90% maximal inhibition). Six other mutations show variable levels of expression but exhibit normal or even gain-of-function (GOF) behavior, as demonstrated by significantly lower EC50 values for α2A adrenergic receptor-mediated inhibition of cAMP. The GNAO1 LOF mutations are associated with epileptic encephalopathy while GOF mutants (such as G42R, G203R and E246K) or normally functioning mutants (R209) were found in patients with movement disorders with or without seizures. Conclusions: Both LOF and GOF mutations in Gαo (encoded by GNAO1) are associated with neurological pathophysiology. There appears to be a strong predictive correlation between the in vitro biochemical phenotype and the clinical pattern of 67! epilepsy vs. movement disorder. 2.2 Introduction Epilepsy is one of the most common neurological disorders in the United States. Severe early onset seizures can result in epileptic encephalopathy (Capovilla, Wolf, Beccaria, & Avanzini, 2013). There are at least 52 different gene mutations that cause early infantile epileptiform encephalopathy (EIEE) (McTague, Howell, Cross, Kurian, & Scheffer, 2016). Mutations in the same genes also cause other neurodevelopmental abnormalities (Berkovic et al., 2004; Sherr, 2003). A key challenge in genetic epilepsies has been understanding the genotype/phenotype relations of causal genes; this may require biochemical analysis (Noebels, 2015). Mutations in the heterotrimeric G protein Gαo (GNAO1 gene) cause an autosomal dominant epileptiform encephalopathy (EIEE17, OMIM: 615473) (Nakamura et al., 2013). In this original paper, all 4 mutations were characterized as having loss-of-function (LOF). More recently, an extended spectrum of “GNAO1 encephalopathies” was identified in which individuals had movement disorders but minimal to no seizures (Ananth et al., 2016; Marce-Grau et al., 2016). Here, we use “GNAO1 encephalopathy” to describe the entire clinical spectrum of individuals with pathological GNAO1 mutations. A genotype-phenotype correlation was also recently noted (Menke et al., 2016); certain mutations (e.g. E246K and several R209 alleles) were found specifically in children with hypotonia, developmental delay, and chorea but no epilepsy. However, the mechanistic 68! basis for this genotype-phenotype correlation remains unknown. GNAO1 encodes the α subunit of Go, a heterotrimeric G protein (consisting of α and βγ subunits) which is highly abundant in the central nervous system, comprising about 1% of brain membrane protein (Sternweis & Robishaw, 1984). Gαo mediates signals from a wide-range of inhibitory receptors including GABAB, α2A adrenergic, adenosine A1 and dopamine D2 receptors. A canonical function of Gαi/o family proteins is inhibition of cAMP (Ghahremani, Cheng, Lembo, & Albert, 1999). The identification of LOF mutations in ADCY5 (which encodes adenylate cyclase 5, the enzyme that produces cAMP) in dyskinesia and chorea patients directly links reduced cAMP to involuntary movement disorders (Carapito et al., 2015; Chang et al., 2016; Mencacci et al., 2015; Morgan, Kurek, Davis, & Sethi, 2016; Raskind et al., 2017). This is inconsistent with LOF behavior of Gαo. Here we assessed expression and function of human mutations in GNAO1 (Allen et al., 2013; Ananth et al., 2016; Consortium, Project, & Consortium, 2014; Dhamija, Mink, Shah, & Goodkin, 2016; Dietel, 2016; Epi KCEaekce, 2016; Kulkarni, Tang, Bhardwaj, Bernes, & Grebe, 2016; Law et al., 2015; Marce-Grau et al., 2016; Nakamura et al., 2013; Saitsu et al., 2016; Talvik, 2015; Zhu et al., 2015). Our biochemical analysis identified both loss- and gain-of-function behaviors; the latter are associated with movement disorders while the former are primarily found in individuals with epileptiform encephalopathies. This mechanistic insight has important implications for therapies of 69! GNAO1 encephalopathies. Appendix B provides updated data on an additional list of 25 mutations analyzed after the publication of Feng et al. 2017. 2.3 Materials and Methods 2.3.1 Materials UK14,304 was from Sigma Aldrich (St Louis, MO). Forskolin was from Calbiochem (San Diego, CA). Pertussis toxin (PTX) was from List Biological Laboratories (Campbell, CA). Protease inhibitor cocktail was from Roche (Roche, Indianapolis, IN). If not otherwise specified, all tissue culture reagents were from Thermo Fisher Scientific (Waltham, MA) and all chemicals were from Sigma Aldrich. 2.3.2 DNA constructs and mutagenesis Cloning of the porcine α2A AR in the pEGFP vector with an amino-terminal HA tag has been previously reported (Brink, Wade, & Neubig, 2000). PTX-insensitive murine Gαo (Gnao1; Gαo p.C351G PTXi) (Jeong & Ikeda, 2000) and RGS- and PTX-insensitive murine Gαo (Gαo p.G184S/C351G; RGS/PTXi) (Jeong & Ikeda, 2000) in the pCI vector were obtained from Dr. Stephen Ikeda (Guthrie Research Institute, Sayre, PA). Using the PTXi murine Gαo C351G as a template, 15 point mutations or deletions were introduced using the Stratagene Quickchange II Site-Directed Mutagenesis kit (Agilent Technologies, Santa Clara, CA). Primers for mutagenesis were designed using the algorithm described at http://www.stratagene.com/sdmdesigner/default.aspx (Table 70! S2.1). Mutations were verified by DNA sequencing at the RTSF Genomics Core at Michigan State University and sequences were analyzed using Clone Manager 9 (Sci-Ed Software, Denver, CO). The protein sequence of the human and murine Gαo are highly similar (98% identical) and do not differ in sequence at any of the mutated positions (Figure S2.1). As a positive control the RGS-insensitive GOF mutant Gαo p.G184S (Lan et al., 1998) was used. 2.3.3 Cell culture and transfections Human embryonic kidney (HEK-293T) cells were maintained in a humidified incubator at 37°C with 5% CO2 and grown to 95% confluence in Dulbecco’s modified Eagle’s medium (DMEM) supplemented with 10% fetal bovine serum (FBS), 100 U/ml penicillin and 100 µg/ml streptomycin. Cells were transfected using Lipofectamine 2000 according to the manufacturer’s recommended protocol. All transfections were performed under serum-free conditions in Opti-MEM. Transfections were allowed to proceed for 4-5 h before the media was changed back to DMEM with 10 % FBS. Experiments were run 24 h after transfection. For western blot, cells in 6-well plates were transfected using 2 µg of DNA and 8 µl of Lipofectamine 2000 per well. For cAMP assays, cells were plated in 60-mm dishes. DNA was kept constant at 4 µg (2 µg of Gαo or pcDNA and 2 µg of α2A AR) and 10 µl of Lipofectamine2000 per plate was used. In the dominant-negative study, a total of 8 µg DNA was added to cells in 60-mm dishes with 10 µl Lipofectamine2000 per plate. In all 71! cases, empty vector (pcDNA3.1) was used to adjust the total amount of DNA. 2.3.4 SDS-PAGE and Western blot Cells were harvested at 4oC in lysis buffer (20 mM Tris-HCl, pH7.4, 150 mM NaCl, 1 mM EDTA, 1 mM β-glycerophospate, 1% Triton X-100, 0.1% SDS, with protease inhibitor) and then sonicated for 10 min at 4oC. Total protein concentrations in the cell lysates were determined by BCA protein assay (Pierce; Rockford, IL) and adjusted with an appropriate volume of Laemmli buffer (BioRad; Hercules, CA) with 5 % 2-mercaptoethanol (β-ME). Equal amounts of protein in each lane were resolved on a 12 % SDS-PAGE gel for 1 h at 160 V. Samples were transferred to an Immobilon-FL PVDF membrane (Millipore, Billerica, MA) for 1 h at 100 V, 400 mA on ice and subjected to Quantitive Infrared Western immunoblot analysis. The membrane was immersed in Odyssey blocking buffer (LI-COR Biosciences, Lincoln, NE) for 1 h with gentle shaking at room temperature. The membrane was simultaneously incubated with anti-Gαo (rabbit; 1:1,000; sc-387; Santa Cruz biotechnologies, Santa Cruz, CA) and anti-actin (goat; 1:1,000; sc-1615; Santa Cruz) antibodies diluted in Odyssey blocking buffer with 0.1% Tween-20 overnight at 4oC. Following four 5 min washes in phosphate-buffered saline, 0.1 % Tween-20 (PBS-T), the membrane was incubated for 1h at room temperature with secondary antibodies (both 1:10,000; IRDye® 800CW Donkey anti-rabbit; IRDye® 680RD Donkey anti-goat; LI-COR Biosciences) diluted in Odyssey blocking buffer with 0.1 % Tween-20. The membrane was subjected to four 5 min washes in PBS-T and a 72! final rinse in PBS for 5 min. The membrane was kept in the dark and the infrared signals at 680 and 800 nm were detected with an Odyssey Fc image system (LI-COR Biosciences). The Gαo polyclonal antibody recognizes an epitope located between positions 90-140 Gαo (Santa Cruz, personal communication), which shouldn’t be affected by any of the mutations studied. 2.3.5 cAMP measurements LANCE Ultra cAMP assays (Perkin Elmer; Waltham, MA) were performed in accordance with the manufacturer’s instructions. Briefly, HEK-293T cells were transfected as indicated above. 100 ng/ml PTX was added the day before the assay to inhibit endogenous Gi/o proteins. Cells were dissociated from dishes using Versene on the day of experiment. Then cells (2,000 cells/well in 5 µl) were transferred to a white 384-well microplate (Perkin Elmer) and incubated with various concentrations of UK14,304 and Forskolin (final 1µM; 5 µl/well) for 30 min at room temperature. A cAMP standard curve was generated in triplicate according to the manual. Finally, europium (Eu)-cAMP tracer (5µL) and ULight™-anti-cAMP (5µL) were added to each well and incubated for 1h at room temperature. The plate was read on a TR-FRET microplate reader (Synergy NEO; Biotek, Winooski, VT). 2.3.6 Data analysis and statistics Quantification of infrared (IR) Western blot signals was performed using Image Studio Lite (LI-COR Biosciences). Individual bands were normalized 73! to the corresponding actin signals, and WT Gαo was set as control. All data was analyzed using GraphPad Prism 6.0 (GraphPad; LaJolla, CA). Dose response curves were fit using non-linear least squares regression. Expression levels and Normalized % inhibition were analyzed with one-way ANOVA with Bonferroni’s post hoc tests for multiple comparisons. Log EC50 values for the NF and GOF mutants were analyzed by paired t-test. Data are presented as mean ± SEM. and a p-value less than 0.05 was considered significant. 2.4 Results 2.4.1 Most pathogenic GNAO1 mutations cause reduced Gαo protein expression To evaluate 15 mutations in GNAO1 (Figure 2.1A and S2.1) that were previously identified in patients with epilepsy or other neurodevelopmental disorders (Allen et al., 2013; Ananth et al., 2016; Consortium et al., 2014; Dietel, 2016; Epi KCEaekce, 2016; Kulkarni et al., 2016; Law et al., 2015; Marce-Grau et al., 2016; Nakamura et al., 2013; Saitsu et al., 2016; Talvik, 2015; Zhu et al., 2015), we performed Western blots in HEK-293T cells transiently transfected with each mutant. The majority of mutants (12) showed significantly lower protein levels than wildtype (WT) Gαo, whereas three separate Arg209 mutant alleles showed essentially normal expression as did the previously described GOF mutant G184S (Fu et al., 2004) (Figure 2.1B-E). 74! Figure 2.1 Location and protein expression levels of human GNAO1 mutations related to epileptic encephalopathy. (A) Location of 15 mutations (G40R, G42R, D174G, T191_F197del, L199P, G203R, R209C, R209G, R209H, A227V, Y231C, E246K, N270H, F275S, and I279N) mapped on the Gαo amino acid sequence. (B, C) Representative Western blots of Gαo protein expression from HEK293T cells transiently transfected with each Gαo mutant. (D, E) Quantification of relative protein levels of each Gαo mutant compared to WT Gαo. Graphs are the result of 3 independent experiments and data are presented as mean ± SEM. p<0.01**, p<0.0001**** using One-way ANOVA with Bonferroni’s post-hoc test for pairwise comparison. 75! 2.4.2 Validation of an in vitro assay to assess function of GNAO1 mutations We used inhibition of forskolin-stimulated cAMP levels as a functional readout to allow efficient quantification of Gαo effects with full concentration curves for agonist-mediated signaling. We co-transfected Gαo plasmids with α2A adrenergic receptor (α2A AR) cDNA (Goldenstein et al., 2009). Robust inhibition of cAMP by the α2 adrenergic agonist UK14,304 depends on both the transfected receptor and the Gαo protein (Figure S2.2A, S2.2B). A pertussis toxin (PTX)-insensitive Gαo (C351G) (Ikeda & Jeong, 2004) was used to create all mutant constructs and WT, enabling inactivation of endogenous Gi/o proteins using PTX. When PTX eliminates the inhibitory signaling through Gi/o, α2A AR couples weakly to Gs and stimulates adenylate cyclase (AC) (Wade et al., 1999), resulting in increased cAMP levels after PTX treatment in the absence of a transfected Gαo (Figure S2.2B). Consequently, the fractional inhibition of AC by Gαo mutants was assessed as the decrease from the high control level of cAMP (PTX but no Gαo - 0%) to the low level with WT Gαo (PTX and PTXi Gαo - 100%). This is termed Normalized % inhibition. PTX treatment did not alter the ability of the PTX-insensitive Gαo to mediate α2A AR-stimulated cAMP inhibition (Figure S2.2C). Hence, this is a good system to study functional consequences of Gαo mutations. 76! 2.4.3 Nine GNAO1 mutations result in loss- (LOF) or partial loss-of-function (PLOF) Six mutants showed essentially complete loss-of-function (LOF) with normalized inhibition below 40% (Figure 2.2A and Table 2.1). All of these mutants also showed low expression levels (11-35% of control; Figure 2.1). Three mutants (A227V, Y231C, and I279N; Figure 2.2A, 2.2E and Table 2.1) had intermediate effects and were classified as partial loss-of-function (PLOF) mutants. The I279N mutation showed a modestly reduced maximal inhibition of cAMP levels (Figure 2.2E and Table 2.1). This mutant also produced a very low EC50 value (0.7 nM vs 25 nM for WT Gαo), which might explain the discrepancy between the quite low expression levels (15% of WT) while maintaining good maximal inhibition in the cAMP inhibition assay. Based on the maximum inhibition below 90% of control, however, we classified this mutation as PLOF (Tables 2.1 and 2.3). The dominant nature of the clinical picture in the GNAO1 encephalopathies raised the question of whether the LOF mutations are actually dominant negative mutations that interfere with the function of the remaining normal Gαo protein expressed in heterozygous individuals. However, in co-expression studies of WT and mutant Gαo at plasmid ratios of 1:1 or 1:2, there was no evidence of a dominant negative action (Figure S2.3). This suggests that the effect of LOF mutations is through a haploinsufficiency mechanism rather than a dominant negative one. 77! Figure 2.2 Effect on α2A AR-mediated cAMP inhibition by GNAO1 mutants. (A-C) Dose-response curves of representative GNAO1 mutants. (A) Dose-response curves of LOF and PLOF mutants (G40R, L199P, N270H, F275S, A227V, Y231C) showing changes in cAMP production in response to the AC activator forskolin and α2AR agonist UK14,304, compared to the positive control (WT) and negative control (pcDNA). (B) Dose-response curves of functioning Gαo mutants showing changes in cAMP production in response to the α2AR agonist UK14,304. All dose-response curves are shown in comparison with WT and G184S. (C) G42R displays a biphasic dose-response 78! Figure 2.2 (cont’d) curve with cAMP inhibition at low concentrations (GOF), followed by enhancement of cAMP levels at higher concentrations of UK14,304. (D) Quantification of EC50 of functioning Gαo mutants. G42R, G203R and E246K exhibit significantly increased potency for α2A AR-mediated cAMP inhibition similar to the known GOF mutation G184S. p<0.05*, p<0.01**, p<0.001*** using paired t test between WT and each mutant separately (Figure S2.4). E. Percentage of maximum inhibition (n=5) was normalized to pcDNA (0%; resulting in activation of cAMP) and WT (100%). p<0.0001**** using One-way ANOVA with Bonferroni’s post-hoc test for pairwise comparison. Note that the maximum inhibition of G42R was calculated at UK14,304 of 15.8 nM. Table 2.1 Functional data for loss-of-function (LOF) and partial loss-of-function (PLOF) mutants. Mutations pcDNA (no Gαo) WT Expression (% of WT)+ cAMP at 5µM UK14,304 (% of un-stimulated)+ Normalized % inhibition+ LogEC50+ EC50 (nM) 0 100 880 ± 30 23 ± 1 0 100 -6.48 ± 0.05 526 -7.71 ± 0.07 25 118G>A Gly40Arg G40R 27 ± 1 **** 900 ± 70 -2 ± 8 **** -6.52 ± 0.13 355 521A>G Asp174Gly D174G 21 ± 7 **** 250 ± 20 -1 ± 7 **** -6.41 ± 0.19 502 517_592del Thr191_Phe197del T191_F197del 11 ± 3 **** 290 ± 20 -16 ± 7 **** -6.36 ± 0.20 525 596T>C Leu199Pro L199P 27 ± 2 **** 570 ± 30 36 ± 3 **** -6.12 ± 0.07 784 680C>T Ala227Val A227V 24 ± 3 **** 320 ± 20 65 ± 3 **** -6.32 ± 0.13 586 808A>C Asp270His N270H 30 ± 2 **** 860 ± 60 2 ± 7 **** -6.42 ± 0.16 499 692A>G Tyr231Cys Y231C 35 ± 2 **** 340 ± 20 63 ± 2 **** -6.34 ± 0.06 472 824T>C Phe275Ser F275S 22 ± 4 **** 720 ± 60 19 ± 7 **** -6.46 ± 0.12 395 836T>A Ile279Asp I279N 15 ± 4 **** 100 ± 10 84 ± 1 -9.18 ± 0.15 0.7 + Mean ± SEM; **, p < 0.01; ***, p < 0.001; ****, p < 0.0001, one-way ANOVA 79! 2.4.4 Six GNAO1 mutations result in gain-of-function (GOF) or normal function (NF) Unexpectedly, a significant number of pathological GNAO1 mutants showed essentially normal or even GOF behavior (Figure 2.2B-E and Table 2.2). As a benchmark for GOF behavior, we used our previously described RGS-insensitive G184S mutant (Fu et al., 2004; Goldenstein et al., 2009; Kehrl et al., 2014; Lan et al., 1998), which shows a mild seizure phenotype in mouse models (Kehrl et al., 2014). It produced a small increase in the maximum inhibition of forskolin-stimulated cAMP levels in response to UK14,304 (Table 2.2). More importantly, it had a significantly more potent response to the α2AR agonist UK14,304 (Figure 2.2D & S2.4A). This represents a 2-3-fold increase in signal strength at low agonist concentrations. Three of the human pathological GNAO1 mutants also showed GOF behavior by this criterion. The G203R, E246K and G42R mutants produced robust inhibition of cAMP with significantly lower EC50 values for the α2AR agonist UK14,304 (Figure 2.2D, S2.4 and Table 2.2). G203R and E246K showed normal inhibition with modest decreases in EC50 (Table 2.2, Figure 2.2D, 2.2E and S.4). This is similar to the effect on EC50 seen for the bona fide GOF mutant G184S. The G42R mutant showed the lowest EC50 of any of the mutants (Figure S2.4), at least 50-fold lower than the WT protein (Figure 2.2C, 2.2D). However, the inhibition mediated by G42R is followed by activation of cAMP with increasing concentrations of UK14,304 (Figure 2.2C). The calculated Normalized % inhibition for 80! the G42R mutant is essentially identical to that of WT Gαo which combined with its very high potency for agonist-mediated inhibition suggests GOF behavior. Three other patient-derived mutations (R209G, R209H, and R209C) showed almost completely normal function and nearly normal expression levels and are designated as normal function (NF) mutants (Figure 2.2B, 2.2D-E, S2.4, and Tables 2.2 & 2.3). The EC50 values for R209G and R209H mutant were not significantly different from WT, while the value for R209C was modestly but significantly higher (Table 2.2, Figure 2.2D & S2.4). 81! Table 2.2 Functional data for normal and gain-of-function (GOF) mutants. Group pcDNA (no Gαo) WT Expression (% of WT)+ cAMP at 5µM UK14,304 (% of unstimulated)+ Normalized % inhibition+ LogEC50+ EC50 (nM) -6.48 ± 0.05 525 -7.71 ± 0.07 0 100 0 100 880 ± 30 23 ± 1 15 ± 1 550G>A++ Gly184Ser G184S 105 ± 5 108 ± 0.1 -8.06 ± 0.09 † 124G>C Gly42Arg G42R 70 ± 6 **** 170 ± 10 99 ± 0.1 -9.34 ± 0.09 ††† 607G>A Gly203Arg G203R 72 ± 9 **** 736G>A Glu246Lys E246K 74 ± 6 ** 625C>G Arg209Gly R209G 77 ± 5 626G>A Arg209His R209H 109 ± 7 625C>T Arg209Cys R209C 96 ± 6 31 ± 3 39 ± 2 74 ± 4 21 ± 1 52 ± 2 100 ± 1.3 -8.06 ± 0.06 ††† 98 ± 0.2 -8.00 ± 0.12 † 96 ± 0.2 -7.75 ± 0.09 100 ± 0.1 -7.52 ± 0.02 97 ± 0.2 -7.39 ± 0.08 †† 25 9.7 0.5 9.3 12 18 30 47 + Mean ± SEM; ++ Not a human mutation; Underlined EC50 Values are significantly different from that of WT; **, p < 0.01; ***, p < 0.001; ****, p < 0.0001, one-way ANOVA; †, p<0.05; ††, p < 0.01; †††, p < 0.001 by paired t-test 82! Table 2.3 Correlation between cAMP inhibition and clinical diagnosis. # Human GNAO1 Mutations 1 124G> Gly42Arg G42R cAMP Inh. GOF C 2 736G>A Glu246Lys E246K GOF 3 736G>A Glu246Lys E246K GOF 4 736G>A Glu246Lys E246K GOF 5 736G>A Glu246Lys E246K GOF 6 736G>A Glu246Lys E246K GOF 7 625C> Arg209Gly R209G NF G 8 626G>A Arg209His R209H NF ! Epilepsy Movement Disorder Age Sex Ref. ++ Unknown F (Zhu et al., 2015) ++ 13 yrs F (Saitsu et al., 2016) ++ 5.5 yrs M (Ananth (twins) et al., 2016) ++ 5.5 yrs F (Ananth (twins) et al., 2016) ++ Decease F (Ananth d at 10 yrs 3 months et al., 2016) ++ 15 yrs M (Ananth et al., 2016) ++ Decease F (Ananth d at 4 yrs 7 months et al., 2016) ++ 16 yrs M (Ananth et al., 2016) 83! Table 2.3 (cont’d) 9 626G>A Arg209His R209H NF 10 626G>A Arg209His R209H NF 11 626G>A Arg209His R209H NF ++ 8 yrs M (Kulkarn i et al., 2016) ++ 6yrs M (Kulkarn i et al., 2016) ++ 5 yrs M (Radhik a Dhamija , 2016) 12 625C>T Arg209Cys R209C NF + ++ 18 yrs F (Saitsu et al., 2016) 13 607G>A Gly203Arg G203R GOF ++ ++ 8 yrs F (Nakam ura et al., 2013) 14 607G>A Gly203Arg G203R GOF + ++ 14 F (Saitsu 15 607G>A Gly203Arg G203R GOF ++ ++ 3 yrs F (Dietel, months et al., 2016) 2016) 4 yrs F (Talvik, 9 months 2015) 20 F (Saitsu months et al., 2016) 16 692A>G Tyr231Cys Y231C PLOF ++ 17 680C>T Ala227Val ! A227V PLOF ++ 84! Table 2.3 (cont’d) 18 836T>A Ile279Asp I279N PLOF ++ 19 836T>A Ile279Asp I279N PLOF ++ 20 118G>A Gly40Arg G40R LOF ++ 21 521A>G Asp174Gly D174G LOF ++ 13 yrs F (Nakam ura et al., 2013) 2 yrs M (Epi KCEaek ce, 2016) 10 F (Law et months al., 2015) 4 yrs F (Nakam 1 month ura et al., 2013) 22 517_59 Thr191_ T191_ LOF ++ + Decease F (Nakam 2del Phe197del F197del d at 11 months ura et al., 2013) 23 596T>C Leu199Pro L199P LOF ++ + 20 F (Marce- 24 824T>C Phe275Ser F275S LOF ++ 25 808A>C Asp270His N270H LOF ++ ! months Grau et al., 2016) 9 yrs F (Allen et al., 2013) 3 yrs F (Consort ium et al., 2014) GOF: gain of function; NF: normal function; PLOF: partial loss of function; LOF: loss of function. ++: major symptoms, +: minor symptoms 85! 2.4.5 Clinical correlation with biochemical behavior of mutant GNAO1 alleles To address genotype/phenotype correlations for GNAO1 encephalopathy, we reviewed the case reports of all 25 individuals who had GNAO1 mutations that had been reported by April, 2016. They have a range of clinical patterns, which extend from early severe epileptic encephalopathy with prominent tonic seizure activity to individuals with a dominant choreo-athetotic movement disorder with virtually no evidence of seizures. There are also individuals (Dietel, 2016; Saitsu et al., 2016), including one of the original 4 cases (Nakamura et al., 2013) (patient #13 – G203R; Table 2.3), who had multiple seizures but also showed prominent choreo-athetosis. In 2016, a clinical report (Ananth et al., 2016) described a unique series of 6 patients with GNAO1 mutations and a pronounced movement disorder, virtually without seizures. They had global developmental delay and hypotonia from infancy and all developed chorea by ages 4-11. In the majority of cases it was intractable, leading to death in two cases. Four patients carried the E246K allele, which we have found to be a GOF mutation. The other two mutations found in this group (R209G and R209C) exhibited essentially normal function in our cAMP inhibition measurements. There are several other reports (Allen et al., 2013; Ananth et al., 2016; Consortium et al., 2014; Epi KCEaekce, 2016; Kulkarni et al., 2016; Law et al., 2015; Marce-Grau et al., 2016; Saitsu et al., 2016; Talvik, 2015; Zhu et al., 2015) of GNAO1 mutations in individuals with a predominant movement disorder with or without seizures. This distinction of clinical 86! patterns based on certain mutant alleles in GNAO1 encephalopathy patients was noted very recently (Menke et al., 2016) but without information about biochemical mechanisms. Table 2.3 summarizes the Gαo biochemical function from the present report and its relation to seizure disorder or movement disorder in literature reports for these mutations. GOF and NF mutations are nearly always found when movement disorder is the predominant feature of the clinical pattern. Mutations that have pure LOF or PLOF biochemical phenotypes are seen in individuals with epileptic encephalopathy without pronounced choreoathetosis. A number of patients exhibit both seizures and movement disorder. We have indicated in Table 2.3 with + or ++ which of these features is predominant or less so. Further studies will be needed based on new cases and/or additional mutations, but there does appear to be a clear pattern emerging about a genotype-phenotype correlation that is driven by a GOF/LOF difference in mutant GNAO1 alleles. 2.4.6 Location of mutations linked to GNAO1 encephalopathies in the Gαo protein To investigate the structural basis for the effects of mutations in Gαo, the mutations were modelled onto the published crystal structure of Gαo in complex with RGS16 (PDB: 3C7K; Figure S2.5). The locations of mutations within the Gαo structure segregated according to their function. The GOF mutations are all near G184S and close to the ribose and phosphate moieties of the bound GDP. The LOF mutants are more broadly scattered throughout the GTPase domain and may destabilize protein folding or stability 87! consistent with their markedly reduced expression levels. The PLOF mutations are clustered in the GTPase domain but away from the bound GDP. This striking structure-function correlation may facilitate prediction of the function of new mutations but ultimately a rigorous biochemical analysis will provide definitive understanding of function. 2.5 Discussion The concept of a “GNAO1 encephalopathy” has developed based on the identification of at least 15 different mutations in the GNAO1 gene (Allen et al., 2013; Ananth et al., 2016; Consortium et al., 2014; Epi KCEaekce, 2016; Kulkarni et al., 2016; Law et al., 2015; Marce-Grau et al., 2016; Nakamura et al., 2013; Saitsu et al., 2016; Talvik, 2015; Zhu et al., 2015) associated with various combinations of epilepsy, developmental delay, hypotonia, and choreo-athetotic movement disorders. Our study demonstrates GOF as well as LOF mutations in GNAO1 and describes a clear correlation between biochemical and clinical characteristics. The existence of these unexpected GOF mutations has important therapeutic implications. Specifically, one might expect that different approaches to therapy would be needed for different mutations (i.e. agonists for LOF and antagonists for GOF mutants). We chose inhibition of cAMP production as the functional readout to assess the Gαo mutants because of the robust measurements permitting complete agonist concentration-response studies. This was critical to our findings since the GOF mutants 88! were detected primarily through their ability to increase signals at low agonist concentrations (Figure 2.2 and Table 2.2). Our previously studied GOF mutant (G184S) which is insensitive to the inhibitory influence of RGS proteins shows such a “left-shift” of agonist concentration response curves in vitro (Clark, Harrison, Zhong, Neubig, & Traynor, 2003; Fu et al., 2004) and in vivo (Goldenstein et al., 2009; Lamberts et al., 2013), and also has a mild seizure phenotype in a mouse model (Kehrl et al., 2014). One might argue that cAMP is not the best choice of functional measures for epilepsy since N-type Ca++ channels or the synaptic release mechanism proteins are critical for regulation of neurotransmitter release. However, the apparent correlation of clinical patterns with the biochemical behavior in our cAMP assay does suggest that function assessed in this way is relevant to functionality in humans. The clear pathological effect of the R209 mutations (with at least 3 individuals carrying distinct alleles), however, does raise the question of why a protein with normal expression and function would cause pathology. It is possible that the R209 mutations have a selective loss of one of the other functional outputs while retaining a normal ability to inhibit AC. Alternatively; there may be selective alterations in expression or localization in neurons that are not accurately reflected in our HEK-293T cell studies of cAMP regulation. A full understanding of the causal mechanisms in GNAO1 encephalopathies requires additional studies of these mutant Gαo proteins in neurons and with different functional readouts. The locations of mutations in the protein structure may partially explain their 89! functional influences. All functioning mutants (NF and GOF) are located around the RGS binding domain, while most of the LOF or PLOF mutants are near the GDP binding region. Two exceptions are D174G and T191_F197del. D174 forms a salt bridge with R162 and mutations in this position may disrupt this interaction. T191_F197del truncates two beta sheets as well as their linking region, which would be expected to decrease protein stability of Gαo. A dominant genetic effect from GOF mutants is not unusual but the fact that the LOF mutations result in a severe autosomal dominant disorder is a bit surprising. We have ruled out a biochemical dominant negative mechanism of these mutations, at least for cAMP regulation, suggesting a haploinsufficiency mechanism. In mice, homozygous Gαo knockouts exhibit seizures as well as hyperactive turning behavior (Jiang et al., 1998). We did not, however, observe spontaneous seizures or an increased sensitivity to pentylenetetrazol (PTZ) kindling in heterozygous Gnao1+/- knockouts (Kehrl et al., 2014). This suggests that humans are more susceptible to haploinsufficiency of Gαo than are mice. In contrast, we observed enhanced kindling sensitivity and reduced survival in our Gnao1+/G184S knock-in mouse model possibly due to seizures (Kehrl et al., 2014). Furthermore, these mice display early neonatal lethality (Kehrl et al., 2014) of unclear mechanism which may be similar to the hypotonia seen in human patients carrying GOF mutations. We do not know whether the abnormalities in these mice are due to brain developmental abnormalities or acute signaling effects. Further studies are needed to 90! better understand this and to determine whether our Gnao1+/G184S mutant mouse might represent a useful pre-clinical model for individuals with GNAO1 GOF mutations. To date, all characterized GNAO1 mutants have been reported as LOF mutations. The G203R mutant in the original paper (Nakamura et al., 2013) was reported as a LOF mutant for regulation of N-type Ca++ channels. Similarly, a G42R mutation in Gαi1 was reported as a LOF mutant based on biochemical studies (Bosch et al., 2012). The unique approach that we have taken with detailed cAMP dose-response studies in a mammalian cell model permitted our recognition of the GOF mechanisms (e.g. the Gαo G42R mutation). The patient with the G203R mutation, which we found to have GOF for cAMP inhibition, had a very different clinical pattern than the other 3 patients in the original study. She had a much later onset of disease (7 months) as well as developmental delay and severe chorea with only localized seizures (Nakamura et al., 2013). A similar clinical pattern was observed in two more, recently described, patients with this same mutation (Dietel, 2016; Saitsu et al., 2016). All patients carrying the GOF mutations identified here appear distinct from the strict EIEE pattern (see Table 3). In comparison, patients with LOF or PLOF mutations were diagnosed with either Ohtahara syndrome (Y231C, I279N, D174G, T191_F197del) or early-infantile epileptic encephalopathy (A227V, L199P, N270H, F275S). Thus GOF and LOF mutations in Gαo appear to result in different disease mechanisms likely requiring different therapeutic approaches. It has remained challenging to convert knowledge about genetic epilepsy mutations 91! into therapies. The GNAO1 encephalopathies may be different because of the eminently targettable nature of the receptors that drive Gαo signaling pathways. A critical question then becomes which receptors might be involved. Interestingly, activation of many Gi/o-coupled receptors is associated with suppression of seizures. Adenosine A1 receptors may play a role in the efficacy of the ketogenic diet (Masino et al., 2011) and agonists at group II metabotropic glutamate receptors are anticonvulsant in various models (Dalby & Thomsen, 1996). The opposite situation is also seen; GABABR agonists exacerbate absence seizures while GABABR antagonists suppress them (Han HA, 2012). Identifying which receptors or downstream signaling effectors of Gαo contribute to mechanisms of encephalopathy from LOF or GOF GNAO1 mutations could therefore reveal potential targets for novel anti-convulsant drug development. We have identified distinct biochemical mechanisms of pathogenic human GNAO1 mutations that may improve the understanding of the heterogeneous clinical spectrum of GNAO1-associated epilepsy and movement disorders. Furthermore, these results also carry significant implications for personalized therapeutics in GNAO1 encephalopathies. 92! APPENDICES 93! APPENDIX A SUPPLEMENTAL DATA 94! Table S2.1 Primer sequences for mutagenesis to create GNAO1 mutants. Mutation Protein R209H (Arg209His) F: 5'-GTGGATCCACTTCTTGTGTTCAGATCGCTGGCC-3' R: 5'-GGCCAGCGATCTGAACACAAGAAGTGGATCCAC-3' Primer Sequences DNA 626 G>A G203R (Gly203Arg) E246K (Glu246Lys) G42R (Gly42Arg) R209C (Arg209Cys) I279N (Ile279Asp) T191_F197del (Thr191_Phe1 97del) R209G (Arg209Gly) A227V (Ala227Val) F275S (Phe275Ser) N270H (Asp270His) G40R (Gly40Arg) 607 G>A 736 G>A F: 5’-AGATCGCTGGCCTCTGACGTCAAACAGCC-3’, R: 5’-GGCTCTTTGACGTCAGAGGCCAGCGATCT-3’ F: 5'-AAGAGCATGAGAGACTTGTGCATGCGGTTCGTG-3' R: 5'-CACGAACCGCATGCACAAGTCTCTCATGCTCTT-3' 124 G>C F: 5'-TTTTTCCTGATTCTCGAGCCCCCAGCAGGAG-3' R: 5'-CTCCTGCTGGGGGCTCGAGAATCAGGAAAAA-3' 625 C>T 836T >A 572_ 592d el F: 5'-GATCCACTTCTTGCATTCAGATCGCTGGCCCC-3' R: 5'-GGGGCCAGCGATCTGAATGCAAGAAGTGGATC-3' F: 5’-GTCAAAGGTGACTTCTTGTTCTTCTCGCCAAAGAGA-3’ R: 5’-ATCTCTTTGGCGAGAAGAACAAGAAGTCACCTTTGAC-3’ F: 5’-GCATCGTAGAAACCCACTTCAGGCTGTTTGACGTC-3’, R:5’-GACGTCAAACAGCCTGAAGTGGGTTTCTACGATGC-3’ 625 C>G Forward 5'-ATCCACTTCTTGCCTTCAGATCGCTGGCCC-3' Reverse 5'-GGGCCAGCGATCTGAAGGCAAGAAGTGGAT-3' 680 C>T F: 5'-GGTCATAGCCGCTGAGTACGACACAGAAGATGATG-3', R: 5'-CATCATCTTCTGTGTCGTACTCAGCGGCTATGACC-3' 824T >C 808A >C 118 G>A F: 5'-ATCTTCTCGCCAGAGAGGTCTTTCTTGTTGAGGAAG-3', R:5'-CATTCCTCAACAAGAAAGACCTCTCTGGCGAGAAGAT-3' F:5'-CAAAGAGGTCTTTCTTGTGGAGGAAGAGGATGATGGA-3 R:5'-TCCATCATCCTCTTCCTCCACAAGAAAGACCTCTTTG-3' F: 5'-CCTGATTCTCCAGCCCTCAGCAGGAGTAATTTC-3' R: 5'- GAAATTACTCCTGCTGAGGGCTGGAGAATCAGG -3' 95! Table S2.1 (cont’d) D174G (Asp174Gly) L199P (Leu199Pro) Y231C (Tyr231Cys) 692A>G 521A>G 596T>C F: 5’-GGTTCGGAGGATGCCCTGCTCGGTGGG-3’, R: 5’-CCCACCGCGCAGGGCATCCTCCGAACC-3’ F: 5'-CCCCGACGTCAAACGGCCTGAAGTGGAGG-3' R: 5'-CCTCCACTTCAGGCCGTTTGACGTCGGGG-3' F: 5’-AGCACCTGGTCACAGCCGCTGAGTGCG-3’ R: 5’-CGCACTCAGCGGCTGTGACCAGGTGCT-3’ 96! Figure S2.1 Alignment of the human and mouse Gαo protein sequences. Human Gαo (NCBI accession number NP_066268.1) and mouse Gαo (NCBI accession number NP_034438.1) were aligned using Clustal Omega(Sievers et al., 2011). The Sequences are 98% identical at the protein level. All the mutations in the current study are highlighted in blue and the inter-species homology is 100% in all those positions. The known gain-of-function mutation G184S is highlighted in green. 97! Figure S2.2 Validation of the Lance Ultra cAMP assay with transient transfection of α2A adrenergic receptor (α2A AR) and Pertussis toxin (PTX)-insensitive Gαo. (A) Co-expression of both α2A AR and Gαo C351G (PTX insensitive) results in strong inhibition of forskolin-stimulated cAMP levels. In the absence of Gαo the α2A AR produces a very modest inhibitory effect. (n=3). (B) α2A AR activation results in cAMP activation in the presence of PTX (n=3). (C) Dose-response curves of co-expression of α2A AR and Gαo C351G exhibit similar EC50 in the absence and presence of PTX (with PTX EC50=9.4nM; without PTX EC50=9.9nM, n=3). 98! Figure S2.3 Assessment of dominant-negative effect of complete LOF mutants G40R, N270H, D174G, T191_F197del, L199P, and F275S. (A) Dose response curves of changes in cAMP production of HEK293T cells co-transfected with different combinations of WT/D174G and α2A AR. D174G did not show an increase of inhibition when co-transfected with WT in different concentrations (n=3). (B) Dose response curves of changes in cAMP production of HEK293T cells co-transfected with different combination of WT/T191_F197del and α2A AR. Although a slight upward shift was observed with WT/T191_F197del 2µg/4µg, the trend did not continue with WT/T191_F197del 2µg/4µg (n=3). (C) Dose response curves of changes 99! Figure S2.3 (cont’d) in cAMP production of HEK293T cells co-transfected with different combination of WT/L199P and α2A AR (n=2). (D) Dose response curves of changes in cAMP production of HEK293T cells co-transfected with different combination of WT/F275S and α2A AR (n=2). (E) Dose response curves of changes in cAMP production of HEK293T cells co-transfected with different combination of WT/G40R and α2A AR (n=2). (F) Dose response curves of changes in cAMP production of HEK293T cells co-transfected with different combinations of WT/N270H and α2A AR (n=2). 100! Figure S2.4 Paired t-test analysis of LogEC50 values for normal and GOF mutants. (A-D) Apart from the known gain of function mutant G184S, G42R, G203R and E246K show a significant decrease of EC50 (n=5). (E) R209C shows an increase of EC50 (n=5). (F, G) R209G and R209H do not display any significant change in EC50 (n=3). 101! Figure S2.5 Mapping of mutations on the structure of Gαo-GDP bound to RGS16. (A) Gαo`GDP ` RGS16 complex with the nucleotide-binding domain in cyan, α-helical domain in grey, switch regions in blue and RGS16 in orange (PDB: 3C7K). (B-D) Localization of mutations on the Gαo`GDP complex. LOF mutants are in red, PLOF mutants are in yellow, GOF mutants are in green and NF mutants are in grey. Each mutant presented in the structure has been changed to its mutated amino acid. (B) R209H serves as a representative to all three mutants at R209. (D) T191_F197del has only been labeled out in red. Protein structure adapted from Slep et al.(Slep et al., 2008) using Pymol (The PyMOL Molecular Graphics System, Version 1.8 Schrödinger, LLC) 102! APPENDIX B VALIDATION OF THE GENOTYPE-PHENOTYPE CORRELATION OF GNAO1-ASSOCIATED NEUROLOGICAL DISORDERS 103! Since the first fifteen GNAO1 mutations covered by our previous study (Feng et al., 2017), many more GNAO1 mutations have been reported. To validate our genotype-phenotype correlation model, we have tested twenty-five newly reported mutations’ functional changes using our cAMP assay. Previously, we used HEK293T cells transiently transfected with α2AR and Gαo proteins for all of our assays. However, in HEK293T cells, α2AR couples to Gαs when the system lacks Gαi or Gαo protein (Figure S2.6C) (Wade et al., 1999). This complicates the interpretation of the results. When HEK293T cells are transfected with LOF mutations or only pcDNA, stimulation of the α2AR results in an increase in cAMP concentration (Figure S2.6C). Here we used another HEK cell line without the Gαs protein (kindly provided by Dr. Kirill Martemyanov from the Scripps Institute; GNAS KO cell line; Figure S2.6) (Masuho et al., 2018), and tested the functions of newly reported GNAO1 mutations. The maximum inhibition of cAMP is less prominent in the GNAS KO cell line (Figure S2.6A & S2.6B). However, in this new system, non-functioning mutations transfected cells, including cells only transfected with pcDNA and α2AR, show no effects on cAMP production (Figure S2.6C, S2.6D & S2.7F, S2.8D). The functioning mutations inhibited the cAMP production when α2AR was activated by UK14,304 (Figure S2.6B, D, F & S2.7E & S2.8C). One interesting aspect using the GNAS KO cell line is that there is a clear rightward shift in the concentration-response curve when the system contains more Gαo protein (Figure S2.6B: without PTX, EC50 for α2AR vs EC50 for α2AR + Gαo: 1.2 nM vs 10.6 nM); S2.6F: EC50 for 104! α2AR + Gαo - PTX vs EC50 for α2AR + Gαo + PTX: 10.6 nM vs 42.6 nM). This rightward shift can be explained by the Gβγ-mediated cAMP inhibition. All adenylate cyclase (AC) isoforms can be stimulated by forskolin, however, stimulated AC activities can be further regulated in a subtype specific manner (Ammer & Christ, 2002). For example, type I AC not only can be inhibited by Gαo, but also by Gβγ directly (Bayewitch et al., 1998a; Taussig, Quarmby, & Gilman, 1993), while type V AC is not affected by Gαo but can be inhibited by Gαi subunits and Gβγ (Bayewitch et al., 1998b; Taussig & Gilman, 1995). Gβγ subunits exhibit a surprising stimulatory effect on type II and type IV ACs, although this effect is highly conditional and only detectable with the presence of GαS (Bayewitch et al., 1998a; Gao & Gilman, 1991; Tang & Gilman, 1991; Taussig & Gilman, 1995). In particular, stimulation of type II AC by Gβγ requires a significantly higher concentration of Gβγ than GαS (Taussig & Gilman, 1995). Human kidney, from which HEK293 cells are developed, mainly expresses Type VI AC (Defer, Best-Belpomme, & Hanoune, 2000), which can be inhibited by Gβγ dimers (Bayewitch et al., 1998b). It is possible that in this system Gαo functions as a restraint for Gβγ’s inhibition of AC; therefore with more Gαo present, the EC50 increases significantly. Similar to our previously described trend, functioning Gαo mutants showed a relatively normal protein expression pattern comparing to WT Gαo (Figure S2.7A, C & S2.8A, B) with the exception of Q223P (Figure S2.7; 0.066 ± 0.019). Non-functioning GNAO1 mutations all exhibited significantly reduced protein expression pattern (Figure 105! S2.7B, D & S2.8A, B). Some of the mutant alleles identified were obtained from genomic database or personal communication from families so clinical information is limited in these cases. However, it is still clear that functioning GNAO1 mutations are associated with movement disorder patients, while non-functioning GNAO1 mutations are mainly related to the onset of epilepsy (Table S2.4). The genotype-phenotype between the mutation functions and the onset of epilepsy or movement disorders still stand with the new HEK cell line (Table S2.4). 106! Figure S2.6 Comparison of validation of the Lance Ultra cAMP assay between HEK293T and GNAS KO HEK293 cells with transient transfection of α2A adrenergic receptor (α2AR) and Pertussis toxin (PTX)-insensitive Gαo. Figure S2.6A, S2.6C, & S2.6E are taken from Figure S2.2. (A, B) Co-expression of both α2A AR and Gαo C351G (PTX insensitive) results in strong inhibition of forskolin-stimulated cAMP levels in both cell lines. In the absence of Gαo the α2A AR produces a very modest inhibitory effect in HEK 293T cells but a strong inhibition effect in GNAS KO cells (n=3). (C, D) α2A AR activation results in cAMP activation in the presence of PTX in HEK293T but no effect in GNAS KO cells due to the lack of Gαs (n=3). (E, F) Dose-response curves of co-expression of α2A AR and Gαo C351G exhibit similar EC50 in the absence and presence of PTX in both cell line (HEK 293T: with PTX EC50=9.4 nM; without PTX EC50=9.9 nM, n=3. GNAS KO: with PTX EC50=42.6 nM; without PTX EC50=10.6 nM, n=3). 107! Figure S2.7 GNAO1 mutations’ functionalities correlate to their protein expression patterns. (A-D) Representative and quantification of relative protein levels of each Gαo mutant compared to wild-type Gαo grouped by functioning mutants (A, C) and non-functioning mutants (B, D). (E, F) Dose-response curves for α2 agonist-mediated inhibition of AC with different GNAO1 mutants. In the GNAS KO HEK cell line, functioning GNAO1 mutants inhibit cAMP production (E) while non-functioning GNAO1 mutants do not show any inhibition of cAMP production (F). Graphs are the results of 3 independent experiments and data are presented as mean ± SEM. *p<0.05, **p<0.01, ****p<0.001 using one-way ANOVA analysis of variance with Bonferroni post-hoc test for pairwise comparison. PTX = pertussis toxin. 108! Figure S2.8 GNAO1 mutations’ functionalities correlate to their protein expression patterns (assays done by Nils Wellhausen with a different group of mutations from those in Figure S2.7). (A) Representative mutations of and (B) quantification of relative protein levels of each Gαo mutant compared to wild-type Gαo. (C, D) Dose-response curves for α2AR regulation of AC with GNAO1 mutants. In GNAS KO HEK cell line, functioning GNAO1 mutants inhibit cAMP production (C) while non-functioning GNAO1 mutants do not show any inhibition of cAMP production (D). Graphs are the results of 3 independent experiments and data are presented as mean ± SEM. *p<0.05, **p<0.01, ****p<0.001 using one-way ANOVA analysis of variance with Bonferroni post-hoc test for pairwise comparison. PTX = pertussis toxin. 109! Table S2.2 All non-functioning GNAO1 mutations tested with GNAS KO cells Expression (% of WT) 0 100 99 ± 3 7 ± 2 4 7 0 5 ± 2 10 12 ± 1 6 ± 3 3 ± 2 -2 ± 5 4 ± 1 - 6 ± 4 cAMP at 5 µM UK14, 304 (% of unstimulated) 116 ± 12 64 ± 4 60 ± 7 110 ± 7 103 ± 4 90 ± 4 102 ± 4 107 ± 6 101 ± 7 120 ± 10 123 ± 13 105 ± 2 108 ± 7 98 ± 5 110 ± 5 86 ± 3 Normalized % Inhibition 0 100 129 ± 14 13 ± 14 29 ± 9 60 ± 9 32 ± 8 21 ± 12 35 ± 14 -10 ± 21 -18 ± 27 -1 ± 4 -10 ± 17 15 ± 12 -31 ± 16 43 ± 7 LogEC50 -5.70 ± 1.30 -7.14 ± 0.67 -7.43 ± 0.07 - - - - - - - - - - - - -6.66 ± 0.17 EC50 (nM) 2014 73 38 - - - - - - - - - - - - 218 Group pcDNA WT G184S Gly184Ser Gly45Gln G45E G45R Gly45Arg Asp273Val D273V Tyr291Asn Y291N Leu157Pro L157P G40W Gly40Trp N270Y Asn270Tyr E246G Glu246Gly L284S Leu284Ser S229R Ser229Arg Leu39Pro L39P Gln52Pro Q52P G40E Gly40Glu 550G>A 134G>A 133G>C 818A>T 871T>A 470T>C 118G>T 808A>T 737A>G 851T>C 687C>A 116T>C 155A>C 119G>A Table S2.3 All functioning GNAO1 mutations tested with GNAS KO cells 550G>A 698A > C 626G>T 844T>A 167T>C 139A>G 709G>A 812A>G 863T>C 725A>C 448T>C 649G>A 1030_1032 delATT Group pcDNA WT G184S Gly184Ser Q233P Gln233Pro R209L Arg209Leu Ser282Thr S282T Ile56Thr I56T Ser47Gly S47G E237K Glu237Lys Lys271Arg K271R Phe288Ser F288S Asn242Thr N242T I163T Ile163Thr Glu217Lys E217K Expression (% of WT) 0 100 106 ± 11 7 ± 2 102 ± 36 63 ± 15 39 ± 6 27 ± 8 24 ± 2 5 ± 1 4 ± 1 - - 65 ± 14 cAMP at 5µM UK14, 304 (% of unstimulated) 116 ± 12 64 ± 4 60 ± 7 73 ± 5 70 ± 4 63 ± 3 79 ± 6 70 ± 6 66 ± 5 97 ± 8 73 ± 4 73 ± 4 80 ± 4 85 ± 5 Normalized % Inhibition 0 100 129 ± 14 100 ± 10 107 ± 7 123 ± 6 87 ± 12 106 ± 12 115 ± 10 17 ± 19 74 ± 10 93 ± 10 37 ± 11 46 ± 12 LogEC50 -5.70 ± 1.30 -7.14 ± 0.67 -7.43 ± 0.07 -7.60 ± 0.32 -6.93 ± 0.19 -7.85 ± 5.08 -7.77 ± 0.31 -7.72 ± 0.29 -7.37 ± 0.17 -8.00 ± 0.88 -7.62 ± 0.31 -7.50 ± 0.24 -7.17 ± 1.12 -6.71 ± 0.64 EC50 (nM) 2014 73 38 25 116 14 17 19 43 10 24 31 67 195 I344del Ile344del 49 ± 15 95 ± 4 22 ± 9 -7.42 ± 0.22 38 110! Table S2.4 Genotype-phenotype correlation of the newly reported GNAO1 mutations GNAO1 Mutations Movement Disorder cAMP Inhibition Epilepsy Gender Age of onset birth 4 y 13 mo 4 mo 3 mo NA 5 mo 12 mo infancy NA 6 mo NA birth 2.5 mo NA NA 15 hr 2 hr 11 d M F F M M F M F F M F F F F M F M F F R209L I56T Q233P E237K E237K S282T* S47G I344del G45E G45R E246G N270Y* G40W* Y291N* L157P* D273V* G40E G40E L284S Functioning Functioning Functioning Functioning Functioning Functioning Functioning Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning Non-Functioning + + + ++ ++ ++ + ++ + ++ + + ++ + ++ ++ ++ + + + ++ + + + + + + * Personal communication from GNAO1 patient support group NA: not available ++ Severe symptoms; + Mild symtoms 111! REFERENCES 112! REFERENCES Allen, A. S., Berkovic, S. F., Cossette, P., Delanty, N., Dlugos, D., Eichler, E. E., . . . Project, E. P. G. (2013). De novo mutations in epileptic encephalopathies. Nature, 501(7466), 217-221. doi:10.1038/nature12439 Ammer, H., & Christ, T. E. (2002). Identity of adenylyl cyclase isoform determines the G protein mediating chronic opioid-induced adenylyl cyclase supersensitivity. J Neurochem, 83(4), 818-827. Ananth, A. L., Robichaux-Viehoever, A., Kim, Y. M., Hanson-Kahn, A., Cox, R., Enns, G. M., . . . Bernstein, J. A. (2016). Clinical Course of Six Children With GNAO1 Mutations Causing a Severe and Distinctive Movement Disorder. Pediatr Neurol. doi:10.1016/j.pediatrneurol.2016.02.018 Bayewitch, M. L., Avidor-Reiss, T., Levy, R., Pfeuffer, T., Nevo, I., Simonds, W. F., & Vogel, Z. (1998a). Differential modulation of adenylyl cyclases I and II by various G beta subunits. J Biol Chem, 273(4), 2273-2276. doi:10.1074/jbc.273.4.2273 Bayewitch, M. L., Avidor-Reiss, T., Levy, R., Pfeuffer, T., Nevo, I., Simonds, W. F., & Vogel, Z. (1998b). Inhibition of adenylyl cyclase isoforms V and VI by various Gbetagamma subunits. 1019-1025. doi:10.1096/fasebj.12.11.1019 FASEB 12(11), J, Berkovic, S. F., Heron, S. E., Giordano, L., Marini, C., Guerrini, R., Kaplan, R. E., . . . Scheffer, I. E. (2004). Benign familial neonatal-infantile seizures: characterization of 550-557. sodium doi:10.1002/ana.20029 channelopathy. Ann Neurol, 55(4), new a Bosch, D. E., Willard, F. S., Ramanujam, R., Kimple, A. J., Willard, M. D., Naqvi, N. I., & Siderovski, D. P. (2012). A P-loop mutation in Gα subunits prevents transition to the active state: implications for G-protein signaling in fungal pathogenesis. PLoS Pathog, 8(2), e1002553. doi:10.1371/journal.ppat.1002553 Brink, C. B., Wade, S. M., & Neubig, R. R. (2000). Agonist-directed trafficking of porcine alpha(2A)-adrenergic receptor signaling in Chinese hamster ovary cells: l-isoproterenol selectively activates G(s). J Pharmacol Exp Ther, 294(2), 539-547. 113! Capovilla, G., Wolf, P., Beccaria, F., & Avanzini, G. (2013). The history of the concept of epileptic encephalopathy. Epilepsia, 54 Suppl 8, 2-5. doi:10.1111/epi.12416 Carapito, R., Paul, N., Untrau, M., Le Gentil, M., Ott, L., Alsaleh, G., . . . Bahram, S. (2015). A de novo ADCY5 mutation causes early-onset autosomal dominant chorea and dystonia. Mov Disord, 30(3), 423-427. doi:10.1002/mds.26115 Chang, F. C., Westenberger, A., Dale, R. C., Smith, M., Pall, H. S., Perez-Duenas, B., . . . Fung, V. S. (2016). Phenotypic insights into ADCY5-associated disease. Mov Disord, 31(7), 1033-1040. doi:10.1002/mds.26598 Clark, M. J., Harrison, C., Zhong, H., Neubig, R. R., & Traynor, J. R. (2003). Endogenous RGS protein action modulates mu-opioid signaling through Galphao. Effects on adenylyl cyclase, extracellular signal-regulated kinases, and intracellular calcium pathways. J Biol Chem, 278(11), 9418-9425. doi:10.1074/jbc.M208885200 Consortium, E.-R., Project, E. P. G., & Consortium, E. K. (2014). De novo mutations in synaptic transmission genes including DNM1 cause epileptic encephalopathies. Am J Hum Genet, 95(4), 360-370. doi:10.1016/j.ajhg.2014.08.013 Dalby, N. O., & Thomsen, C. (1996). Modulation of seizure activity in mice by metabotropic glutamate receptor ligands. J Pharmacol Exp Ther, 276(2), 516-522. Defer, N., Best-Belpomme, M., & Hanoune, J. (2000). Tissue specificity and physiological relevance of various isoforms of adenylyl cyclase. Am J Physiol Renal Physiol, 279(3), F400-416. doi:10.1152/ajprenal.2000.279.3.F400 Dhamija, R., Mink, J. W., Shah, B. B., & Goodkin, H. P. (2016). GNAO1-Associated 615-617. Disorder. Mov Pract, Movement doi:10.1002/mdc3.12344 Disord Clin 3(6), Dietel, T. (2016). Genotype and phenotype in GNAO1-Mutation – Case report of an unusual course of a childhood epilepsy. In G. S. T. B. Haack , I. Cordes , R. Trollmann , T. Bast (Ed.), (Vol. 47, pp. P01-17). Neuropediatrics: Thieme. Epi KCEaekce, E. K. (2016). De Novo Mutations in SLC1A2 and CACNA1A Are Important Causes of Epileptic Encephalopathies. Am J Hum Genet. doi:10.1016/j.ajhg.2016.06.003 114! Feng, H., Sjogren, B., Karaj, B., Shaw, V., Gezer, A., & Neubig, R. R. (2017). Movement disorder in GNAO1 encephalopathy associated with gain-of-function mutations. Neurology, 89(8), 762-770. doi:10.1212/WNL.0000000000004262 Fu, Y., Zhong, H., Nanamori, M., Mortensen, R. M., Huang, X., Lan, K., & Neubig, R. R. (2004). RGS-insensitive G-protein mutations to study the role of endogenous RGS 229-243. Methods doi:10.1016/S0076-6879(04)89014-1 Enzymol, proteins. 389, Gao, B. N., & Gilman, A. G. (1991). Cloning and expression of a widely distributed (type IV) adenylyl cyclase. Proc Natl Acad Sci U S A, 88(22), 10178-10182. doi:10.1073/pnas.88.22.10178 Ghahremani, M. H., Cheng, P., Lembo, P. M., & Albert, P. R. (1999). Distinct roles for Galphai2, Galphai3, and Gbeta gamma in modulation offorskolin- or Gs-mediated cAMP accumulation and calcium mobilization by dopamine D2S receptors. J Biol Chem, 274(14), 9238-9245. Goldenstein, B. L., Nelson, B. W., Xu, K., Luger, E. J., Pribula, J. A., Wald, J. M., . . . Doze, V. A. (2009). Regulator of G protein signaling protein suppression of Galphao protein-mediated alpha2A adrenergic receptor inhibition of mouse hippocampal CA3 epileptiform activity. Mol Pharmacol, 75(5), 1222-1230. doi:mol.108.054296 [pii]10.1124/mol.108.054296 Han HA, C. M., Snead OC III. (2012). GABAB Receptor and Absence Epilepsy. (4th edition. ed.). Jasper's Basic Mechanisms of the Epilepsies [Internet]. Bethesda (MD): National Center for Biotechnology Information (US). Ikeda, S. R., & Jeong, S. W. (2004). Use of RGS-insensitive Galpha subunits to study endogenous RGS protein action on G-protein modulation of N-type calcium channels in sympathetic neurons. Methods Enzymol, 389, 170-189. doi:10.1016/S0076-6879(04)89011-6 Jeong, S. W., & Ikeda, S. R. (2000). Effect of G protein heterotrimer composition on coupling of neurotransmitter receptors to N-type Ca(2+) channel modulation in sympathetic neurons. Proc Natl Acad Sci U S A, 97(2), 907-912. Jiang, M., Gold, M. S., Boulay, G., Spicher, K., Peyton, M., Brabet, P., . . . Birnbaumer, L. (1998). Multiple neurological abnormalities in mice deficient in the G protein Go. Proc Natl Acad Sci U S A, 95(6), 3269-3274. doi:10.1073/pnas.95.6.3269 115! Kehrl, J. M., Sahaya, K., Dalton, H. M., Charbeneau, R. A., Kohut, K. T., Gilbert, K., . . . Neubig, R. R. (2014). Gain-of-function mutation in Gnao1: a murine model of epileptiform encephalopathy (EIEE17)? Mamm Genome, 25(5-6), 202-210. doi:10.1007/s00335-014-9509-z Kulkarni, N., Tang, S., Bhardwaj, R., Bernes, S., & Grebe, T. A. (2016). Progressive Movement Disorder in Brothers Carrying a GNAO1 Mutation Responsive to Deep Brain 211-214. doi:10.1177/0883073815587945 Stimulation. 31(2), J Child Neurol, Lamberts, J. T., Smith, C. E., Li, M. H., Ingram, S. L., Neubig, R. R., & Traynor, J. R. (2013). Differential control of opioid antinociception to thermal stimuli in a knock-in mouse expressing regulator of G-protein signaling-insensitive Galphao protein. J Neurosci, 33(10), 4369-4377. doi:10.1523/JNEUROSCI.5470-12.2013 Lan, K. L., Sarvazyan, N. A., Taussig, R., Mackenzie, R. G., DiBello, P. R., Dohlman, H. G., & Neubig, R. R. (1998). A point mutation in Galphao and Galphai1 blocks interaction with regulator of G protein signaling proteins. J Biol Chem, 273(21), 12794-12797. Law, C. Y., Chang, S. T., Cho, S. Y., Yau, E. K., Ng, G. S., Fong, N. C., & Lam, C. W. (2015). Clinical whole-exome sequencing reveals a novel missense pathogenic variant of GNAO1 in a patient with infantile-onset epilepsy. Clin Chim Acta, 451(Pt B), 292-296. doi:10.1016/j.cca.2015.10.011 Marce-Grau, A., Dalton, J., Lopez-Pison, J., Garcia-Jimenez, M. C., Monge-Galindo, L., Cuenca-Leon, E., . . . Macaya, A. (2016). GNAO1 encephalopathy: further delineation of a severe neurodevelopmental syndrome affecting females. Orphanet J Rare Dis, 11, 38. doi:10.1186/s13023-016-0416-0 Masino, S. A., Li, T., Theofilas, P., Sandau, U. S., Ruskin, D. N., Fredholm, B. B., . . . Boison, D. (2011). A ketogenic diet suppresses seizures in mice through adenosine A⁾ receptors. J Clin Invest, 121(7), 2679-2683. doi:10.1172/JCI57813 Masuho, I., Chavali, S., Muntean, B. S., Skamangas, N. K., Simonyan, K., Patil, D. N., . . . Martemyanov, K. A. (2018). Molecular Deconvolution Platform to Establish Disease Mechanisms by Surveying GPCR Signaling. Cell Rep, 24(3), 557-568 e555. doi:10.1016/j.celrep.2018.06.080 McTague, A., Howell, K. B., Cross, J. H., Kurian, M. A., & Scheffer, I. E. (2016). The 116! genetic landscape of the epileptic encephalopathies of infancy and childhood. Lancet Neurol, 15(3), 304-316. doi:10.1016/S1474-4422(15)00250-1 Mencacci, N. E., Erro, R., Wiethoff, S., Hersheson, J., Ryten, M., Balint, B., . . . Bhatia, K. P. (2015). ADCY5 mutations are another cause of benign hereditary chorea. Neurology, 85(1), 80-88. doi:10.1212/WNL.0000000000001720 Menke, L. A., Engelen, M., Alders, M., Odekerken, V. J., Baas, F., & Cobben, J. M. (2016). Recurrent GNAO1 Mutations Associated With Developmental Delay and a Movement Disorder. J Child Neurol. doi:10.1177/0883073816666474 Morgan, J. C., Kurek, J. A., Davis, J., & Sethi, K. D. (2016). ADCY5 mutations are another cause of benign hereditary chorea. Neurology, 86(10), 978-979. doi:10.1212/WNL.0000000000002479 Nakamura, K., Kodera, H., Akita, T., Shiina, M., Kato, M., Hoshino, H., . . . Saitsu, H. (2013). De Novo mutations in GNAO1, encoding a Gαo subunit of heterotrimeric G proteins, cause epileptic encephalopathy. Am J Hum Genet, 93(3), 496-505. doi:10.1016/j.ajhg.2013.07.014 Noebels, J. (2015). Pathway-driven discovery of epilepsy genes. Nat Neurosci, 18(3), 344-350. doi:10.1038/nn.3933 Raskind, W. H., Friedman, J. R., Roze, E., Meneret, A., Chen, D. H., & Bird, T. D. (2017). ADCY5-related dyskinesia: Comments on characteristic manifestations and variant-associated severity. Mov Disord, 32(2), 305-306. doi:10.1002/mds.26888 Saitsu, H., Fukai, R., Ben-Zeev, B., Sakai, Y., Mimaki, M., Okamoto, N., . . . Matsumoto, N. (2016). Phenotypic spectrum of GNAO1 variants: epileptic encephalopathy to involuntary movements with severe developmental delay. Eur J Hum Genet, 24(1), 129-134. doi:10.1038/ejhg.2015.92 Sherr, E. H. (2003). The ARX story (epilepsy, mental retardation, autism, and cerebral malformations): one gene leads to many phenotypes. Curr Opin Pediatr, 15(6), 567-571. Sievers, F., Wilm, A., Dineen, D., Gibson, T. J., Karplus, K., Li, W., . . . Higgins, D. G. (2011). Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol Syst Biol, 7, 539. doi:10.1038/msb.2011.75 117! Slep, K. C., Kercher, M. A., Wieland, T., Chen, C. K., Simon, M. I., & Sigler, P. B. (2008). Molecular architecture of Galphao and the structural basis for RGS16-mediated deactivation. Proc Natl Acad Sci U S A, 105(17), 6243-6248. doi:10.1073/pnas.0801569105 Sternweis, P. C., & Robishaw, J. D. (1984). Isolation of two proteins with high affinity for guanine nucleotides from membranes of bovine brain. J Biol Chem, 259(22), 13806-13813. Talvik, I. (2015). Clinical Phenotype of De Novo GNAO1 Mutation: Case Report and Review of Literature. In M. V. Rikke S. Møller, Ulvi Vaher, Line HG Larsen, Hans A. Dahl, Pilvi Ilves, and Tiina Talvik (Ed.), (Vol. 2). Child Neurology Open. Tang, W. J., & Gilman, A. G. (1991). Type-specific regulation of adenylyl cyclase by G 1500-1503. 254(5037), subunits. Science, protein doi:10.1126/science.1962211 gamma beta Taussig, R., & Gilman, A. G. (1995). Mammalian membrane-bound adenylyl cyclases. J Biol Chem, 270(1), 1-4. doi:10.1074/jbc.270.1.1 Taussig, R., Quarmby, L. M., & Gilman, A. G. (1993). Regulation of purified type I and type II adenylylcyclases by G protein beta gamma subunits. J Biol Chem, 268(1), 9-12. Wade, S. M., Lim, W. K., Lan, K. L., Chung, D. A., Nanamori, M., & Neubig, R. R. (1999). G(i) activator region of alpha(2A)-adrenergic receptors: distinct basic residues mediate G(i) versus G(s) activation. Mol Pharmacol, 56(5), 1005-1013. Zhu, X., Petrovski, S., Xie, P., Ruzzo, E. K., Lu, Y. F., McSweeney, K. M., . . . Goldstein, D. B. (2015). Whole-exome sequencing in undiagnosed genetic diseases: interpreting 119 trios. Genet Med, 17(10), 774-781. doi:10.1038/gim.2014.191 118! CHAPTER 3: BEHAVIORAL ASSESSMENT OF MOUSE MODELS WITH GNAO1-ASSOCIATED MOVEMENT DISORDER AND EPILEPSY Modified from Feng, H., Larrivee, C. L., Demireva, E. Y., Xie, H., Leipprandt, J. R., & Neubig, R. R. (2019). Mouse models of GNAO1-associated movement disorder: Allele-and sex-specific differences in phenotypes. PloS one, 14(1), e0211066. DOI:10.1371/journal.pone.0211066 With permission from PLOS ONE. All rights reserved. & From Larrivee, C. L., Feng, H., Leipprandt, J. R., Demireva, E. Y., Xie, H., & Neubig, R. R. (2019). Mice with GNAO1 R209H Movement Disorder Variant Display Hyperlocomotion Alleviated by Risperidone. bioRxiv, 662031. DOI: https://doi.org/10.1101/662031 Demireva, E. Y. and Xie, H. did the design and production of the G203R and R209H mutant mice. Larrivee C. L. performed Rotarod and Open Field studies for the G203R and R209H mutant mice. Leipprandt, J. R. did the breeding and genotyping for the mutant mice. 119! 3.1 Abstract Infants and children with dominant de novo mutations in GNAO1 exhibit movement disorders, epilepsy, or both. Children with loss-of-function (LOF) mutations and partial-loss-of-function (PLOF) mutations exhibit Epileptiform Encephalopathy 17 (EIEE17). Gain-of-function (GOF) or those with normal functioning (NF) mutations in an in vitro assay are found in patients with Neurodevelopmental Disorder with Involuntary Movements (NEDIM). There is no animal model with a human mutant GNAO1 allele. Here we assess the behavioral patterns in several mouse models. Mouse models with Gnao1 knock-in mutation G203R (GOF), R209H (NF) or ΔT191F197 (LOF) were created by CRISPR/Cas9 methods to determine whether the clinical features of patients with a particular GNAO1 mutation which could include epilepsy and/or movement disorder would be evident in the mouse model. These three newly developed models are compared with previously developed Gnao1 mouse model, Gnao1+/G184S and Gnao1+/-. Gnao1+/G203R mutant mice were viable and gained weight comparably to controls. Homozygotes were not non-viable. Grip strength was decreased in both males and females. Male Gnao1+/G203R mice were strongly affected in movement assays (RotaRod and DigiGait) while females were not. Male Gnao1+/G203R mice also showed enhanced seizure propensity in the pentylenetetrazole kindling test. Mice with a G184S GOF knock-in also showed movement-related behavioral phenotypes but females were more strongly affected than males. In contrast, the Gnao1+/R209H mouse model exhibited 120! hyperkinetic movements, which have not been seen in any other Gnao1 mutant mouse line. Gnao1+/R209H mice also did not show a strong sex difference in our behavioral battery test. Mice carrying the strong epilepsy allele, Gnao1+/ΔT191F197 did not gain weight like their WT siblings and the mutant mice developed seizures around P7 and all mutant mice died before P16. In contrast, Gnao1+/- mice survived and never developed spontaneous seizures. Gnao1+/G203R, Gnao1+/R209H and Gnao1+/ΔT191F197 mice all shared similar phenotypes regarding to the onset of epilepsy and/or movement disorders as children with the same heterozygous GNAO1 mutations. Although Gnao1+/ΔT191F197 mice did not survive to breeding age and the line was lost, both Gnao1+/G203R and Gnao1+/R209H mouse models should be useful tools in mechanistic and preclinical studies of GNAO1-related movement disorders and epilepsy. 3.2 Introduction Neurodevelopmental Disorder with Involuntary Movements (NEDIM) is a newly defined neurological disorder associated with mutations in GNAO1. It is characterized by “hypotonia, delayed psychomotor development, and infantile or childhood onset of hyperkinetic involuntary movements” (OMIM 617493). NEDIM is monogenetic and associated with GOF mutations and NF mutations in GNAO1 (Feng et al., 2017). The GNAO1 gene has also been associated with early infantile epileptic encephalopathy 17 (EIEE17; OMIM 615473). However, 36% of patients showed both epilepsy and 121! movement disorder phenotypes. This includes many different GNAO1 alleles such as: G40R, G45R, S47G, I56T, T191_F197del, L199P, G203R, R209C, A227V, Y231C and E246G (Feng, Khalil, Neubig, & Sidiropoulos, 2018). GNAO1 encodes Gαo, the most abundant membrane protein in the mammalian central nervous system (Jiang & Bajpayee, 2009). Gαo is the α-subunit of the Go protein, a member of the Gi/o family of heterotrimeric G proteins. Gi/o proteins couple to many important G protein-coupled-receptors (GPCRs) involved in movement control like GABAB, dopamine D2, adenosine A1 and adrenergic α2A receptors (Franek et al., 1999; Gazi, Nickolls, & Strange, 2003; Lorenzen, Lang, & Schwabe, 1998; Tian, Duzic, Lanier, & Deth, 1994). Upon activation, Gαo and Gβγ separate from each other and modulate separate downstream signaling pathways. Gαo mediates inhibition of cyclic AMP (cAMP), and Gβγ mediates inhibition of AC and N-type calcium channels while activating G-protein activated inward rectifying potassium channels (GIRK channels) (Zhang, Pacheco, & Doupnik, 2002). Go is expressed mainly in the central nervous system and it regulates neurotransmitter release by modulating intracellular calcium concentrations in pre-synaptic cells (Li et al., 2004). It has also been suggested that Go plays a role in neurodevelopmental processes like neurite outgrowth and axon guidance (Bromberg, Iyengar, & He, 2008; Strittmatter, Fishman, & Zhu, 1994). Consequently, Go is an important modulator of neurological functions. Previously, we defined a functional genotype-phenotype correlation for GNAO1 122! disorders (Feng et al., 2017); GOF and NF mutations are found in patients with movement disorders, while loss-of-function (LOF) and partial-loss-of-function (PLOF) mutations are associated with epilepsy (Feng et al., 2017). I recently published a mechanistic review of this genotype-phenotype correlation (Feng et al., 2018). The experimental study of human GNAO1 mutant alleles, however, was done in HEK293T cells, which lack complex physiological context of the brain. Therefore, it was important to see whether mouse models with GNAO1 mutations would share clinical characteristics of the human patients. Such a result would verify the previously - reported genotype-phenotype correlation and would provide a system for more detailed mechanistic studies and preclinical testing models for possible new therapeutics. Previously, we reported that Gnao1+/G184S mutant mice carrying a human-engineered GOF mutation (G184S) showed heightened sensitization to pentylenetetrazol (PTZ) kindling and had an elevated frequency of interictal epileptiform discharges on EEG (Kehrl et al., 2014). Here, we tested whether the Gnao1+/G184S mice also exhibit movement disorders although G184S has not been found in human (Feng et al., 2019). A Gnao1+/- mouse model was also described previously. They are hyperalgesic and display severe motor control impairment (Jiang et al., 1998). Gnao1-/- mice are hyperactive and also exhibit an abnormal turning behavior (Jiang et al., 1998). In our hands, they are poorly viable. The G203R is a GOF GNAO1 mutation in the cAMP assay (Feng et al., 2017). It is 123! one of the most common GNAO1 mutations found clinically (Table 3.1) (Arya, Spaeth, Gilbert, Leach, & Holland, 2017; Feng et al., 2018; Nakamura et al., 2013; Saitsu et al., 2016; Schorling et al., 2017; Xiong et al., 2018). Most patients with this mutation exhibit both seizures and movement disorders (Arya et al., 2017; Feng et al., 2018; Nakamura et al., 2013; Saitsu et al., 2016; Schorling et al., 2017; Xiong et al., 2018). The R209 is a mutation hotspot in the Gαo protein, which has been reported in over ten patients. Previously, the R209H mutation was in more than seven patients (Ananth et al., 2016; R. Dhamija, Mink, Shah, & Goodkin, 2016; Kelly et al., 2019; Kulkarni, Tang, Bhardwaj, Bernes, & Grebe, 2016; Marecos, Duarte, Alonso, Calado, & Moreira, 2018), all of who develop severe chorea/athetosis, dystonia, hypotonia and developmental delay (Table 3.1). All three previously reported R209 mutations (R209H, R209C, and R209G) were NF GNAO1 mutations in our cAMP assay (Feng et al., 2017). However, it remained a question why GNAO1 mutations that showed normal function in the cellular assay were pathogenic. The ΔT191F197 in-frame deletion mutation was the most severe LOF pathogenic GNAO1 mutation reported in patients (Feng et al., 2017; Nakamura et al., 2013) (Table 3.1). It is the most representative LOF mutation with the lowest protein expression level and lowest % inhibition (Feng et al., 2017); therefore it was a good candidate for generating a LOF mouse model that would possibly develop spontaneous neurological disorders. However, since all mutant ΔT191F197 mice died before P16, we were not 124! able to generate a line carrying this mutation. In this chapter, I only provide limited data on survival and growth of this mouse model. Table 3.1 The status of Gnao1 mutant mice Allele cAMP Clinical Mouse Status Inhibition GOF LOF GOF G184S KO G203R N/A N/A EIEE17 & MD Rare adult lethality; breeds well (Kehrl et al., 2014) Homozygous early lethality (Jiang et al., 1998) Rare adult lethality; breeds well (Feng et al., 2019) R209H ΔT191F197 NF LOF MD Hyperactivity; gait phenotypes (Larrivee et al., 2019) EIEE17 Heterozygous early lethality (P7-P14); loss of strains (Figure 3.4) We intended to develop mouse models with the representative human GOF (G203R), NF (R209H), and LOF (ΔT191F197) mutations to see if they replicated the clinical phenotype of GNAO1 mutation-associated neurological disorders. If so, they would be valuable tools to understand neural mechanisms underlying the complex phenotypic spectrum of patients with GNAO1 mutations. In this chapter, I show the behavioral assessment of two mouse lines carrying Gαo GOF mutation Gnao1+/G203R and NF mutation Gnao1+/R209H. They are compared with two previously described mouse models: one with a known GOF function mutation G184S (Gnao1+/G184S) and the other the Gnao1 KO model (Gnao1+/-). These two mouse models (Gnao1+/G203R and Gnao1+/R209H) present 125! the possibility of studying GNAO1-associated neurological defects in animal models. 3.3 Materials and Methods 3.3.1 Animals Animal studies were performed in accordance with the Guide for the Care and Use of Laboratory Animals established by the National Institutes of Health. All experimental protocols and personnel were approved by the Michigan State University Institutional Animal Care and Use Committee (IACUC). Mice were housed on a 12-h light/dark cycle and had free access to food and water. They were studied between 8-12 weeks old. 3.3.2 Generation of Gnao1 mutant mice 3.3.2.1 Generation of Gnao1+/G203R mouse model Gnao1+/G184S (Feng et al., 2017; Fu et al., 2004; Goldenstein et al., 2009; Kehrl et al., 2014) and Gnao1+/- mice (Kehrl et al., 2014) were generated as previously described and used as N10 or greater backcross on the C57BL/6J background. Gnao1G203R mutant mice were generated using CRISPR/Cas9 genome editing on the C57BL/6NCrl strain. gRNA targets within exon 6 of the Gnao1 locus (ENSMUSG00000031748) were used to generate the G203R mutation (Figure 3.1A). Synthetic single-stranded DNA oligonucleotides (ssODN) were used as repair templates carrying the desired mutation and short homology arms (Table 3.2). CRISPR reagents were delivered as ribonucleoprotein (RNP) complexes. RNPs were assembled in vitro using wild-type S.p. Cas9 Nuclease 3NLS protein, and synthetic tracrRNA and crRNA 126! (Integrated DNA Technologies, Inc.). TracrRNA and crRNA were denatured at 95°C for 5 min and cooled to room temperature in order to form RNA hybrids, which were incubated with Cas9 protein for 5 min at 37°C. RNPs and ssODN templates were electroporated into C57BL/6NCrl zygotes as described previously(Qin et al., 2015), using a Genome Editor electroporator (GEB15, BEX CO, LTD). C57BL/6NCrl embryos were implanted into pseudo-pregnant foster dams. Founders were genotyped by PCR (Table 3.2) followed by T7 Endonuclease I assay (M0302, New England BioLabs) and validated by Sanger sequencing. Figure 3.1 Development of Gnao1+/G203R mouse model. (A) Targeting of the Gnao1 locus. The location of the gRNA target protospacer and the PAM, and double stranded breaks following Cas9 cleavage are indicated on the WT allele. Deleted or modified sequences are highlighted in blue. The resulting edited allele sequence and translation 127! Figure 3.1 (cont’d) are presented along with the sequences used as references for ssODN synthesis. (B) Heterozygous Gnao1+/G203R mutant mice are largely normal in size and behavior. Photo comparing mutant mouse with its littermate control is shown. (C) Gnao1+/G203R mice have a relatively normal survival; while homozygous Gnao1G203R/G203R mice die perinatally (P0-P1). (D) Gnao1+/G203R mice develop normally and gain weight similarly to their WT littermate controls. Table 1. Location, sequence and genotyping of gRNA targets in Gnao1 locus. Table 3.2 Location, sequence and genotyping of gRNA targets in the Gnao1 locus. DSB location gRNA target ssODN Gnao1 G203R chr 8: 93,950,314 5’ TGCAGGCTGTTTGACGTCGG GGG 3’ (+) 5’ ATGGCCGTGACATCCTCAAAGCAGTGGATCCAC TTCTTGCGTTCAGATCGCTGGCCGCGGACGTCAAA CAGTTTGCAGGGAGTCAGGGAAAGCTGT 3’ PCR primers Genotyping Fwd: 5’ GACAGGTGTCACAGGGGATG 3’ Rev: 5’ TCCTAGCCAAGACCCCAACT 3’ PCR product = 462bp SacII site created by G203R mutation gRNA target – 20bp protospacer and PAM sequences are listed, strand orientation indicated by (+) or (-). Sequence of ssODN used as repair template is listed. For G203R, mutated codon is highlighted in bold. DSB – double gRNA target – 20bp protospacer and PAM sequences are listed, strand orientation indicated by (+) or (-). The sequence of the ssODN used as a repair template is listed. For G203R, the mutated codon is highlighted in bold. DSB – double stranded break. PAM – protospacer adjacent motif. stranded break. PAM – protospacer adjacent motif.! The likelihood of an off-target site being edited is very low. Based on the number and position of mismatches several predictive algorithms were used to assign guide 128! specificity scores from 0 to 100 (100 is the best) to rank gRNAs by specificity with respect to off-targets occurring (Doench et al., 2016; Haeussler M, 2016; Hsu et al., 2013). The gRNA target used for this experiment has a specificity score of 94, which is the highest seen by the MSU Transgenic and Gene Editing Facility in over 40 similar targeting experiments (E. Demivera personal communication). This greatly reduces the probability of an off-target edits occurring. After examining the off-target lists (Table S3.9), we did not identify any off-target loci with less than 3 mismatches or with an off-target binding score > 0.5 which we deem as thresholds for further validation. We also did not identify any off-target loci with significant scores that were on the same chromosome and would be less likely to be removed from the genome after breeding of several generations. Furthermore, the RNP (ribonucleoprotein) approach of delivering CRISPR reagents to mouse embryos we employed further lowers the risk of off-target events (Iyer et al., 2018). Nevertheless, we directly validated several predicted off-target loci within coding regions for the G203 gRNA target TGCAGGCTGTTTGACGTCGG GGG. One off-target site with 4 mismatches and a score of 0.52 was validated for locus ENSMUSG00000041390. We also analyzed two other off-target sites with 4 mismatches and scores of 0.15 and 0.069 respectively, predicted to occur on the same chromosome (chr 8) ENSMUSG00000086805 and ENSMUSG00000097637. To test these 3 off-target sites, DNA from WT and founder animals from which the line was expanded were 129! analyzed by PCR and sequencing and we found that no off-target effects had occurred for all 3 off-target loci analyzed (see Supplemental Materials). 3.3.2.2 Generation of Gnao1+/R209H mouse model Mutant Gnao1+/R209H mice were generated via CRISPR/Cas9 genome editing on a C57BL/6J genomic background. CRISPR gRNA selection and locus analysis were performed using the Benchling platform (Benchling, Inc. San Francisco, CA.). A gRNA targeting exon 6 of the Gnao1 locus (ENSMUSG00000031748) was chosen to cause a double strand break (DSB) 3bp downstream of codon R209. A single-stranded oligodeoxynucleotide (ssODN) carrying the R209H mutation CGC > CAC with short homology arms was used as a repair template (Figure 3.2 and Table 3.3). Ribonucleoprotein (RNP) complexes consisting of a synthetic crRNA/tracrRNA hybrid and Alt-R® S.p. Cas9 Nuclease V3 protein (Integrated DNA Technologies, Inc. Coralville, IA), were used to deliver CRISPR components along with the ssODN to mouse zygotes via electroporation as previously described (Feng et al., 2019). Edited embryos were implanted into pseudo-pregnant dams using standard techniques. Resulting litters were screened by PCR (Phire Green HSII PCR Mastermix, F126L, Thermo Fisher, Waltham, MA.), T7 Endonuclease I assay (M0302, New England Biolabs Inc.) and Sanger sequencing (GENEWIZ, Inc. Plainfield, NJ) for edits of the target site. 130! Table 3.3 Location and sequence of gRNA and ssODN template for CRISPR-Cas targeting Gnao1 locus; primers and genotyping method for Gnao1+/R209H mice Location gRNA target 5’ N20-PAM -3’ Gnao1 R209H Chr 8: 93,950,334 5’ AGCGATCTGAACGCAAGAAG TGG 3’ ssODN template (reverse complement) GTTTCGTCCTCGTGGAGCACCTGGTCATAGCCGCT GAGTGCGACACAGAAGATGATGGCCGTGACATCCTCAAA GCAGTGGATCCACTTCTTGtGTTCAGATCGCTGGCCCCCG ACGTCAAACAGCCTGCAGGGAGTCAGGGAAAGCTGTGA GGGCGGGGACGCCTA PCR primers Genotyping O586 FWD: 5' GGACAGGTGTCACAGGGGAT 3’ O587 REV: 5' ACTGGCCTCCCTTGGCAATA 3' By Sanger Sequencing Figure 3.2 Targeting of the mouse Gnao1 locus. (A) Mouse Gnao1 genomic locus (exon size not to scale), red outline is magnified in (B) showing exon 6 and relative location of codon 209, and PCR primers O586 and O587. (C) Location and exact sequence of gRNA target within exon 6, dotted red line denotes DSB, PAM is highlighted 131! Figure 3.2 (cont’d) and sequence corresponding to gRNA protospacer is underlined (also in E). (D) Raw gel electrophoresis images showing PCR of the target region and T7 Endonuclease I (T7 Endo I) digestion analysis of founders 1324 – 1335 (n=12), with WT, H2O (-) and T7Endo I (+) controls. Founder 1324 (red number) was positive for the mutation on one allele and WT on the other, note that the single bp mismatch was not reliably detected by T7 Endo I assay. (E) Exact sequence of edited founder 1324 as aligned to WT reference genome, two peaks (G and A) are detected on the sequence chromatogram, indicating the presence of both WT and edited R209H allele. 3.3.3 Genotyping and Breeding 3.3.3.1 Genotyping Gnao1+/G203R mice Heterozygous Gnao1+/G203R mutant founder mice were crossed against C57BL/6J mice to generate Gnao1+/G203R heterozygotes (N1 backcross). Further breeding was done to produce N2 backcross heterozygotes while male and female N1 heterozygotes were crossed to produce homozygous Gnao1G203R/G203R mutants. Studies were done on N1 or N2 G203R heterozygotes with comparisons to littermate controls. All mice had ears clipped before weaning. DNA was extracted from earclips by an alkaline lysis method (Truett et al., 2000). The G203R allele of Gαo was identified by Sac II digests (WT 462 Bp and G203R 320 & 140Bp) of genomic PCR products generated with primers (Fwd 5' GACAGGTGTCACAGGGGATG 3'; Rev 5' TCCTAGCCAAGACCCCAACT 3'). Reaction conditions were: 0.8 µl template, 4 µl 5x Promega PCR buffer, 0.4 µl 10 mM dNTPs, 1 µl 10 µM Forward Primer, 1 µl 10 µM Reverse Primer, 0.2 µl Promega GoTaq and 12.6 µl DNase free water (Promega catalog 132! # M3005, Madison WI). Samples were denatured for 4 minutes at 95 oC then underwent 32 cycles of PCR (95 oC for 30 seconds, 60 oC for 30 seconds, and 72 oC for 30 seconds) followed by a final extension (7 minutes at 72 oC). After PCR, samples were incubated with Sac II restriction enzyme for 2 hrs. 3.3.3.2 Genotyping Gnao1+/R209H mice Studies were done on N1 R209H heterozygotes with comparisons to littermate controls. To generate Gnao1+/R209H heterozygotes (N1 backcross), 2 founder Gnao1+/R209H mice, 1 male and 1 female, were crossed with C57BL/6J mice. DNA was extracted by an alkaline method (Hirata, Takahashi, Shimoda, & Koide, 2016) from ear clips done before weaning. PCR products were generated with primers flanking the mutation site (Fwd 5' GGACAGGTGTCACAGGGGAT 3’; 5' ACTGGCCTCCCTTGGCAATA 3'). Reaction conditions were: 0.8 µl template, 4 µl 5x Promega PCR buffer, 0.4 µl 10mM dNTPs, 1 µl 10 µM Forward Primer, 1 µl 10 µ M Reverse Primer, 0.2 µl Promega GoTaq and 12.6 µl DNase free water (Promega catalog # M3005, Madison WI). Samples were denatured for 4 minutes at 95 ̊ C then underwent 32 cycles of PCR (95 ̊ C for 30 seconds, 63 ̊C for 30 seconds, and 72 ̊C for 30 seconds) followed by a 7-minute final extension at 72 ̊C. Ethanol precipitation was done on the PCR products and then samples were sent for Sanger sequencing (GENEWIZ, Inc. Plainfield, NJ). 133! 3.3.4 Behavioral Studies Researchers conducting behavioral experiments were blinded until the data analysis was completed. Before each experiment, mice were acclimated in the testing room for at least 10 min. The timeline of behavioral protocols is described in Figure 3.3. Female experimenters conducted all behavioral studies. Figure 3.3 The timeline for utilizing animals in this study. Open field, Rotarod and Grip strength tests were performed on the same group of 8-week-old animals in this as showed above. DigiGait tests were done on naïve 8-week-old animals. Animals finishing the motor behavior studies were used for the PTZ kindling study for 3 weeks. 3.3.4.1 Open Field The Open Field test was conducted in Fusion VersaMax 42 cm x 42 cm x 30 cm arenas (Omnitech Electronics, Inc., Columbus, OH). Mice and their littermate controls were placed in the arena for 30 minutes to observe spontaneous activities. Using the Fusion Software, distance traveled (cm) was evaluated for novel (first 10 minutes), 134! sustained (10-30 minutes), and total (0-30 minutes) activity. Center Time was also measured. Center Time was defined as the time spent in the center portion (20.32cm x 20.32cm) of the Open Field cage. 3.3.4.2 RotaRod Motor skills were assessed using an Economex accelerating RotaRod (Columbus Instruments, Columbus, OH). The entire training and testing protocol took two days. On day 1, mice were trained for three 2-minute sessions, with a 10-minute rest between each training period. During the first two sessions, the RotaRod was maintained at a constant speed of 5 rpm. In the third training session, the rod was started at 5 rpm and accelerated at 0.1 rpm/sec for 2 minutes. On day 2, mice were trained with two more accelerating sessions for 2 minutes each with a 10-minute break in between. The final test session was 5 minutes long, starting at 5 rpm then accelerating to 35 rpm (0.1 rpm/sec). For all training and test trials, the time to fall off the rod was recorded. RotaRod learning curves were done on a separate group of mice with 10 tests in one day with a 5-min rest between each test. The learning rate of each group of animals was calculated as described (Hirata et al., 2016). 3.3.4.3 Grip Strength Mouse grip strength data was collected following a protocol adapted from Deacon et al (Deacon, 2013) using seven home-made weights (10, 18, 26, 34, 42, 49, 57 grams). Briefly, the mouse was held by the middle/base of the tail and lowered to grasp a weight. 135! A total of three seconds was allowed for the mouse to hold the weight with its forepaws and to lift the weight until it was clear of the bench. Three trials were done starting with the 10g weight to permit the mice to lift the weights with a 10-second rest between each trial. If the mouse successfully held a weight for 3 seconds, the next heavier weight was given; otherwise the maximum time/weight achieved was recorded. A final total score was calculated based on the heaviest weight the mouse was able to lift up and the time that it held it (Deacon, 2013). The final score was normalized to the body weight of each mouse, which was measured before the trial. 3.3.4.4 DigiGait Mouse gait data were collected using a DigiGait Imaging System (Mouse Specifics, Inc., Framingham, MA) (Hansen & Pulst, 2013). The test is used for assessment of locomotion as well as the integrity of the cerebellum and muscle tone/equilibrium(Franco-Pons, Torrente, Colomina, & Vilella, 2007). Briefly, after acclimation, mice were allowed to walk on a motorized transparent treadmill belt. A high-speed video camera was mounted below to capture the paw prints on the belt. Each paw image was treated as a paw area and its position recorded relative to the belt. Seven speeds (18, 20, 22, 25, 28, 32 and 36 cm/s) were tested per animal with a 5-minute rest between each speed. An average of 4-6 s of video was saved for each mouse, which is sufficient for the analysis of gait behaviors in mice (Franco-Pons et al., 2007). For each speed, left & right paws were averaged for each animal while fore and 136! hind paws were evaluated separately. Stride length was normalized to animal body length. 3.3.5 PTZ Kindling Susceptibility A PTZ kindling protocol was performed as described before (Kehrl et al., 2014) to assess epileptogenesis. Briefly, PTZ (40 mg/kg, i.p. in 5 mg/ml) was administered every other day starting at 8 weeks of age. Mice were monitored and scored for 30 minutes for behavioral signs of seizures as described(Grecksch et al., 2004; Kehrl et al., 2014; Wilczynski et al., 2008). Kindling is defined as death or the onset of a tonic-clonic seizure on two consecutive treatment days. The number of injections for each mouse to reach a sensitization was reported in survival curves. This experiment lasted up to 4 weeks with a maximum of 12 doses. Each animal in the study was checked every day for health and seizure development. Animals were humanely euthanized with CO2 immediately after kindling or after 12 PTZ injections and observation. In total, 40 animals were used for this study, among which 27 died of tonic-clonic seizures and 13 were euthanized after 12 doses of PTZ injections. 3.3.6 Data Analysis All data was analyzed using GraphPad Prism 7.0 (GraphPad; La Jolla, CA). Data are presented as mean ± SEM and a p value less than 0.05 was considered significant. All statistical tests are detailed in Figure Legends. Multiple comparison correction of the 137! dataset from DigiGait was performed via a false discovery rate (FDR) correction at a threshold value of 0.01 in an R environment using the psych package. 3.4 Results 3.4.1 The growth patterns of the three newly developed Gnao1 mutant mouse models (G203R, R209H and ΔT191F197) 3.4.1.1 Gnao1+/G203R mice showed normal viability and growth. Genotypes of offspring of Gnao1+/G203R x WT crosses (N1 - C57BL/6NCrl x C57BL/6J) were observed at the expected frequency (29 WT and 27 heterozygous). All three homozygous mice from Gnao1+/G203R x Gnao1+/G203R crosses died by P1. The small numbers of offspring observed from these crosses so far, however, were not significantly different from expected frequencies (4 wt, 14 het, and 3 homozygous). Heterozygous Gnao1+/G203R mice did not show any growth abnormalities compared to Gnao1+/+ mice (Figure 3.1B & 3.1D) and they had relatively normal survival. There were two spontaneous deaths (~5-7 weeks) seen for Gnao1+/G203R mice out of 33 (Figure 3.1C). This is reminiscent of the spontaneous deaths seen previously with the Gnao1+/G184S GOF mutant mice (Kehrl et al., 2014). Gnao1+/G203R mice did not exhibit any obvious spontaneous seizures or abnormal movements. 3.4.1.2 Gnao1+/R209H mice have expected frequency and normal viability Two founder Gnao1+/R209H mice, 1 male and 1 female, were crossed with C57BL/6J mice. Out of 98 offspring of a cross of Gnao1+/R209H with WT mice, 51 heterozygotes and 138! 47 WT were observed. Gnao1+/R209H mice exhibit no overt postural or movement abnormalities or seizures at basal conditions. Adult mice showed no statistically significant differences in weight between WT and Gnao1+/R209H genotypes of either sex (data not shown). 3.4.1.3 Gnao1+/ΔT191F197 mice developed spontaneous seizures at P7 and died before P16. Gnao1ΔT191F197 mutant mice were generated using CRISPR/Cas9 genome editing on the C57BL/6NCrl strain. gRNA targets within exon 6 of the Gnao1 locus (ENSMUSG00000031748) were used to generate the ΔT191F197 mutation (Figure 3.4A). Only one viable founder (male) was obtained. Genotypes of offsprings of this male founder Gnao1+/ΔT191F197 x WT crosses (C57BL/6NCrl x C57BL/6J) were observed. Heterozygous Gnao1+/ΔT191F197 mice were very rare and all died perinatally within P16 (Figure 3.4C &3.2D). Gnao1+/ΔT191F197 mice also developed spontaneous seizure at P7 (Figure 3.4B). Previously, we described that Gnao1 G184S heterozygous mutant mice on a 129 background (N6 129S1/SvImJ) lived a relatively normal life (Kehrl et al., 2014). To investigate whether a 129 background is also protective towards the ΔT191F197 mutant mice, we crossed our Gnao1+/ΔT191F197 with WT mice on the 129 background and assessed their offsprings. Unfortunately, 129 alleles did not appear to provide a dominant protective effect against the spontaneous death observed in the heterozygous ΔT191F197 mutant mice (Figure 3.4C). 139! Figure 3.4 Gnao1+/ΔT191F197 mice developed spontaneous seizures at P7 and died before P16. (A) Targeting of the Gnao1 locus. The location of the gRNA target protospacer and the PAM, and double stranded breaks following Cas9 cleavage are indicated on the WT allele. Deleted or modified sequences are highlighted in blue. The resulting edited allele sequence and translation are presented along with the sequences used as references for ssODN synthesis. (B) Video snapshot of one heterozygous Gnao1+/ΔT191F197 mutant mouse developed spontaneous seizure at P7. Photo comparing mutant mouse with its littermate control is shown. (C) Gnao1+/ΔT191F197 on both C57BL/6J and B6/129 backgrounds died prematurely. (D) Gnao1+/ΔT191F197 mice did not develop or gain weight normally comparing to WT littermate controls. 140! 3.4.2 Behavioral assessment of mutant Gnao1 mouse models (G184S, G203R, R209H and KO) for movement patterns 3.4.2.1 Female Gnao1+/G184S and male Gnao1+/G203R mice show similar movement abnormalities and gait disturbances Since GOF alleles of GNAO1 in children result primarily in movement disorder, we tested motor coordination in two mouse lines with GOF mutations. One carried an engineered GOF mutant G184S, designed to block RGS protein binding (DiBello et al., 1998; Fu et al., 2004; Lan et al., 1998). The other is the G203R GOF mutant, which has been seen in at least 9 children (Chapter 1) (Feng et al., 2018; Feng et al., 2017). First, we used a two-day training and testing procedure on the RotaRod (Figure 3.5A & 3.5B). Gnao1+/G184S and Gnao1+/G203R mice were compared to their same-sex littermate controls. Female Gnao1+/G184S mice exhibited a reduced retention time on the accelerating RotaRod (unpaired t-test, p<0.001, Figure 3.5A) while male mice remained unaffected. In contrast, male Gnao1+/G203R mice exhibited reduced time to stay on the rotating rod (unpaired t-test, p<0.05, Figure 3.5B) while female Gnao1+/G203R mice did not show any abnormalities. Results from all the RotaRod training and testing sessions are shown in Figure S1. Neither Gnao1+/G184S nor Gnao1+/G203R mice showed a significant difference in learning rate on RotaRod (Figure S3.3), suggesting that the differences we observed in the RotaRod study was due to movement deficits rather than learning difficulties. Grip strength was assessed as described (Deacon, 2013). This test is widely done in 141! combination with the RotaRod motor coordination test. This may be relevant to the hypotonia, seen in many GNAO1 patients (Ananth et al., 2016; Bruun et al., 2018; Danti et al., 2017; Euro, Epilepsy Phenome/Genome, & Epi, 2014; Gawlinski et al., 2016; Honey et al., 2018; Kulkarni et al., 2016; Law et al., 2015; Marce-Grau et al., 2016; Saitsu et al., 2016; Schorling et al., 2017; Waak et al., 2018; Yilmaz et al., 2016; Zhu et al., 2015). Similar to the RotaRod results, female Gnao1+/G184S mice also showed reduced forepaw grip strength compared to their littermate controls (unpaired student t-test, p<0.05, Figure 3.5C) while males did not exhibit a significant difference (Figure 3.5C). In contrast, both male and female Gnao1+/G203R mice displayed reduced forepaw grip strength (unpaired t-test, p<0.05, Figure 3.5D). 142! Figure 3.5 Female Gnao1+/G184S mice and male Gnao1+/G203R mice show reduced time on RotaRod and reduced grip strength. (A&B) Quantification of RotaRod studies. (A) Female Gnao1+/G184S mice lose the ability to stay on a RotaRod (unpaired t-test; ***p<0.001), while male Gnao1+/G184S mice appeared unaffected. (B) Male Gnao1+/G203R also showed reduced motor coordination on RotaRod (unpaired t-test, *p<0.01). (C&D) Quantification of grip strength results. Scores for each mouse were normalized to the body weight of the animal measured. (C) Female Gnao1+/G184S mice are less capable of lifting weights compared to their Gnao1+/+ siblings (unpaired t-test, *p<0.05). (D) Both male and female Gnao1+/G203R mice showed reduced ability to hold weights (unpaired t-test, *p<0.05). Data are shown as mean ± SEM. 143! The open field test provides simultaneous measurements of locomotion, exploration and surrogates of anxiety. It is a useful tool to assess locomotive impairment in rodents (Tatem et al., 2014), however, environmental salience may reduce the impact of the motor impairment on behaviors (Parr & Friston, 2017). Therefore, we divided the 30-min open field measurements into two periods, with the first 10 min assessing activity in a novel environment and the 10-30 minute period designated as sustained activity (Figure 3.6C & 3.6D). The novelty measurement showed a significant difference between Gnao1+/G184S mice and their littermate controls for both male and female mice (2-way ANOVA, p<0.01 for female, p<0.05 for male, Figure 3.6C). Female, but not male, Gnao1+/G184S mice showed reduced activity in the sustained phase of open field testing (2-way ANOVA, *p<0.05, **p<0.01, ****p<0.0001). Both male and female Gnao1+/G184S mice also showed reduced total activity (2-way ANOVA, p<0.001, Figure 3.6A & 3.6C). Neither male nor female Gnao1+/G203R mice performed differently in the open field arena compared to their littermate controls (Figure 3.6B & 3.6D). No significant difference was observed in the time mice spent in the center of the arena (Figure S3.2). 144! Figure 3.6 G184S mutant mice showed reduced activities in Open Field Test but G203R mutants do not. (A&C) Female and male Gnao1+/G184S mice showed decreased activity in the open field test. A total of 30 min activity was recorded which was divided into Novelty (0-10 min) and Sustained (10-30 min) period. (A) Representative heat map of overall activity comparison between Gnao1+/+ and Gnao1+/G184S mice in both sexes. (C) Quantitatively, both male and female Gnao1+/G184S travelled less in the open field arena (2-way ANOVA; ****p< 0.0001, **p<0.01, *p<0.05). (B & D) Neither male nor female Gnao1+/G203R mice showed abnormalities in the open field arena. (B) Sample heat map tracing both female and male mouse movement in open field. (D) Quantification showed no difference between Gnao1+/+ and Gnao1+/G203R mice in distance traveled (cm) in the open field arena (2-way ANOVA; n.s.). Data are shown as mean ± SEM. Numbers of animals are indicated on bars. 145! In addition to the above behavioral tests, we also performed gait assessment on Gnao1+/G184S and Gnao1+/G203R mice of both sexes. Gait is frequently perturbed in rodent models of human movement disorders even when the actual movement behavior seen in the animals does not precisely phenocopy the clinical movement pattern (Song, Fan, Exeter, Hess, & Jinnah, 2012; Stroobants, Gantois, Pooters, & D'Hooge, 2013). The multiple parameters assessed in DigiGait allow it to pick up subtle neuromotor defects and makes it more informative than the RotaRod test. The gait analysis largely confirmed the sex differences between the two strains in RotaRod tests. Thirty-seven parameters were measured for both front and hind limbs. Given the large number of measurements, we used false discovery rate (FDR) analysis with a Q of 1% as described in Methods to reduce the probability of Type I errors (Figure S3.4 & S3.5, Table S3.1-S3.4). Gnao1+/G184S female mice showed 22 significant differences (Q<0.01) and males showed 8 (Figure S3.4, Table S3.3 & S3.4). For Gnao1+/G203R mice, the opposite sex pattern was seen with 27 parameters in females and 8 parameters in males showing significant differences from WT (Figure S3.5, Table S3.1 & S3.2). Two of the most highly significant parameters and ones that had face validity in terms of clinical observations (stride length and paw angle variability) were chosen for further analysis. Across the range of treadmill speeds, female Gnao1+/G184S mice showed significantly reduced stride length (2-way ANOVA, p<0.01, Figure 3.7A) and increased paw angle 146! variability (2-way ANOVA, p<0.0001, Figure 3.7E) compared to WT littermates. Male Gnao1+/G184S mice only had a difference in paw angle variability (2-way ANOVA, p<0.0001), not in stride length (Figure 3.7C & 3.7G). These results are consistent with the results of RotaRod and grip strength measurements in that female Gnao1+/G184S mice showed a stronger phenotype than males. In contrast to the Gnao1+/G184S mice, male Gnao1+/G203R mice appeared to be more severely affected in gait compared to female Gnao1+/G203R mice. Male Gnao1+/G203R mice had highly significantly reduced stride length (2-way ANOVA, p<0.0001, Figure 3.7D) and increased paw angle variability (2-way ANOVA, p<0.05, Figure 3.7H). In contrast, female Gnao1+/G203R mice did not show any significant differences in stride length or paw angle variability (Figure 3.7B & 3.7F). In addition to these quantitative gait abnormalities a qualitative defect was seen. A significant number of Gnao1+/G203R mice of both sexes failed to run when the belt speed exceeded 22 cm/s (Mann-Whitney test, female and male p<0.05, Figure 3.7J). For reasons that are not clear such a difference was not seen for Gnao1+/G184S mice (Figure 3.7I). 147! Figure 3.7 DigiGait Imaging System reveals sex-specific gait abnormalities in Gnao1+/G184S mice and Gnao1+/G203R mice. (A-D) Female Gnao1+/G184S mice showed significant gait abnormalities, while female Gnao1+/G203R mice remain normal. (A & B) Female Gnao1+/G184S mice showed reduced stride length (2-way ANOVA with Bonferroni multiple comparison post-test) while female Gnao1+/G203R mice were unchanged from control (2-way ANOVA; n.s.). (C) Female Gnao1+/G184S mice also showed increased paw angle variability (2-way ANOVA, p<0.0001) while female Gnao1+/G203R mice showed normal paw angle variability. (E-H) Male Gnao1+/G203R and Gnao1+/G184S mutant mice showed distinct gait abnormalities. (E & G) Male Gnao1+/G184S mice showed significantly 148! Figure 3.7 (cont’d) increased paw angle variability (2-way ANOVA p <0.0001 overall with significant Bonferroni multiple comparison tests; **p<0.01 and *p<0.05). There was no effect on stride length. (F & H) In contrast, male Gnao1+/G203R mice showed markedly reduced stride length (2-way ANOVA p<0.0001 with Bonferroni multiple comparison post-test; ***p<0.001, **p<0.01, and *p<0.05) and modestly elevated paw angle variability (overall p<0.05). (I) Gnao1+/G184S mice did not show significant differences in the highest treadmill speed successfully achieved. (J) Both male and female Gnao1+/G203R mice showed reduced capabilities to run on a treadmill at speeds greater than 25 cm/s (Mann-Whitney test; *p<0.05). 3.4.2.2 Gnao1+/R209H mouse model exhibits unique hyperactive behavior in the open field arena but no abnormal motor coordination in other behavior tests and minor disturbance in gait analysis All patients with R209H mutation were diagnosed with hyperkinetic movements including chorea/athetosis and dystonia (Ananth et al., 2016; R. Dhamija et al., 2016; Kelly et al., 2019; Kulkarni et al., 2016; Marecos et al., 2018). To assess whether Gnao1+/R209H mice phenocopy the patients’ symptoms, we repeated the above behavior tests with the newly developed heterozygous R209H mice. Unlike Gnao1+/G184S and Gnao1+/G203R mice, both sexes of Gnao1+/R209H mice were hyperactive in the open field arena (Figure 3.8A& 3.8B) but completely normal in Rotarod (Figure 3.8C) and grip strength assessment (Figure 3.8D). Male and female Gnao1+/R209H mice also showed reduced time spent in the center of the arena (Figure 3.8A & 3.8B), which is an indication for possible anxiety-linked behavior (Wilmshurst, 149! Byrne, & Webb-Peploe, 1989). Figure 3.8 Gnao1 +/R209H mice show significant hyperactivity and reduced time in center in the open field arena. (A) Representative heat maps of Gnao1+/R209H mice and Gnao1+/+ mice in the open field arena (B) Time spent in the open field arena was separated into 0-10 minutes (novelty) and 10-30 minutes (sustained). Gnao1+/R209H male and female mice exhibit increased locomotion in the novelty period. Hyperactivity was maintained throughout the sustained period as mice continued to show significant 150! Figure 3.8 (cont’d) increase in distance traveled (2-way ANOVA; ****p < 0.0001, ***p < 0.001, * p < 0.05). Gnao1+/R209H mice of both sexes spend less time in center areas of the open field arena compared to WT littermates. (C) Neither male nor female Gnao1+/R209H mice show significant differences on the Rotarod. (D) There is no significant difference between grip strength between WT and Gnao1+/R209H mice. Data are shown as mean ± SEM. A h t g n e L e d i r t S ) h t g n e l y d o b f o % ( C y t i l i l b a i r a V e g n A w a P ) D S ( 90 80 70 60 50 40 15 15 10 5 0 15 F Gnao1+/+ (14) F Gnao1+/R209H (9) 20 25 30 35 Speed (cm/s) F Gnao1+/+ (14) F Gnao1+/R209H (9) B h t g n e L e d i r t S ) h t g n e l y d o b f o % ( 40 D y t i l i b a i r a V e l g n A w a P ) D S ( 25 30 35 Speed (cm/s) 40 40 ** 90 80 70 60 50 40 15 15 10 5 0 15 M Gnao1+/+ (9) M Gnao1+/R209H (10) ** 2-way ANOVA P<0.0001 35 40 20 25 30 Speed (cm/s) M Gnao1+/+ (9) M Gnao1+/R209H (10) 20 25 30 35 40 Speed (cm/s) 20 E ) s / m c ( d e e p S x a M 30 20 10 + / + 1 o a F G n 1 o a F G n H 9 0 2 + / R + / + 1 o a M G n 1 o a M G n H 9 0 2 + / R Figure 3.9 Male and female Gnao1+/R209H mice shows gait abnormalities in different tests on the DigiGait imaging system. (A & B) Male Gnao1+/R209H mice showed reduced stride length compared to wildtype littermates (2-way ANOVA with Bonferroni multiple comparison post-test), while female Gnao1+/R209H mice show a normal stride 151! Figure 3.9 (cont’d) length. (C & D) Neither male nor female Gnao1+/R209H exhibited significant differences in paw angle variability compared to WT littermates. (E) At speeds greater than 25 cm/s female Gnao1+/R209H shows reduced ability to run on a treadmill. Gait analysis was done with a Digigait video system. Similar to Gnao1+/G203R mouse model, male Gnao1+/R209H mice showed a highly significant genotype effect with reduced stride length compared to wildtype littermates (**p<0.01, 2-way ANOVA). Females showed no difference from WT. No difference was seen in paw angle variability of Gnao1+/R209H of either sex. However, both male and female Gnao1+/R209H mice failed to consistently run at higher speeds (>20 cm/s; Figure 3.9E, **p<0.01, Student’s t-test). The difference observed was not due to a reduced body length (WT: 9.54 cm vs R209H: 10.17 cm) or weight. Comparisons of other parameters assessed by the Digigait system were shown in the appendix (Figure S3.6 & Table S3.5-S3.6). Compared to the number of parameters with significant difference detected for Gnao1+/G203R and Gnao1+/G184S mice respectively, Gnao1+/R209H mice only did not display as many significantly different gait abnormalities (Figure S3.6). 3.4.2.3 Previously described Gnao1+/- mouse model did not show any abnormalities in the behavioral test battery A Gnao1+/- KO mouse model was previously described for the study of the mechanisms of Go protein (Jiang et al., 1998; Valenzuela et al., 1997). In these reports, homozygous Gnao1-/- mice lived but exhibited a reduced lifespan and developed severe 152! motor control impairment (Jiang et al., 1998). Gnao1-/- mice were also reported to be hyperactive and had an abnormal turning behavior (Jiang et al., 1998). Heterozygous Gnao1+/- mice did not have any spontaneous abnormal behavior (Jiang et al., 1998). We have been unable to generate homozygous (Gnao1-/-) KO mice. In our behavior battery, heterozygous Gnao1+/- mice did not show any abnormalities in open field (Figure 3.10A-D), Rotarod (Figure 3.10E), or grip strength (Figure 3.10F). However, male Gnao1+/- mice did exhibit significantly reduced stride length (Figure 3.11B, p<0.001, 2-way ANOVA). Furthermore, male Gnao1+/- mice also showed several DigiGait parameters that are significantly different compared to the WT mice (Figure S3.7 & Table S3.7-S3.8). To our surprise, Gnao1+/- mice also did not develop any spontaneous seizure activity in contrast to the LOF mutant mice Gnao1+/ΔT191F197, which were severely impaired by spontaneous seizures perinatally and died prematurely (Figure 3.4). 153! Figure 3.10 Male and female Gnao1+/- mice do not show any abnormalities in the behavioral tests including open field, Rotarod, and grip strength. (A-D) Neither sex of Gnao1+/- mice has normal activity in the open field arena in any time period. The activity pattern of Gnao1+/- mice (A) is also similar to Gnao1+/+ (B). The overall activity is also comparable between male (D) and female (C) WT and Gnao1+/- mice. (E) Gnao1+/- mice do not show any reduced motor coordination capability in Rotarod. (F) Grip strength test shows that neither sex of Gnao1+/- mice decreases their capability of lifting heavy weight. 154! Figure 3.11 DigiGait Imaging System reveals the decreased stride length in male Gnao1+/- mice. Female Gnao1+/- mice do not show any difference comparing to female Gnao1+/- mice in either stride length (A) or paw angle variability (C). (B) However, male Gnao1+/- mice exhibit a significantly decreased stride length (p<0.0001, 2-way ANOVA) but not (D) paw angle variability. 3.4.3 PTZ kindling study of G203R and R209H mice. Kindling studies for Gnao1+/G184S and Gnao1+/- mice have been reported previously by Kehrl et al (Kehrl et al., 2014). 3.4.3.1 Male Gnao1+/G203R mice are sensitized to PTZ kindling. Epilepsy has been observed in 100% of patients with GNAO1 G203R mutations (Arya et al., 2017; Feng et al., 2018; Nakamura et al., 2013; Saitsu et al., 2016; Xiong et al., 2018). Also in the Gnao1+/G184S GOF mutant mice, we previously reported spontaneous lethality as well as increased susceptibility to kindling by the chemical 155! anticonvulsant PTZ for both males and females (Kehrl et al., 2014). Kindling is a phenomenon where a sub-convulsive stimulus, when applied repetitively and intermittently, leads to the generation of full-blown convulsions. To determine if the G203R GOF mutant mice mimicked the G184S mutants and phenocopied the human epilepsy pattern of children with the G203R mutation, we assessed PTZ-induced kindling in Gnao1+/G203R mutant mice. As expected for C57BL/6 mice, females were more prone to kindling than male mice, half kindled at 4 and 8-10 injections, respectively (Figure 3.12A & 3.12B). Despite the increased sensitivity of females in general, female Gnao1+/G203R mice did not show significantly higher sensitivity to PTZ compared to their littermate controls (Figure 3.12A). On the contrary, male Gnao1+/G203R mice were more sensitive to PTZ kindling than controls (Figure 3.12B, Mantel-Cox Test, p<0.05). Also, three spontaneous deaths were seen (two male and one female) among the 33 G203R mice observed for at least 100 days, similar to the early lethality seen in G184S mutant mice. We cannot, however, attribute those deaths to seizures at this point. 156! Figure 3.12 Gnao1+/G203R male mice have an enhanced Pentylenetetrazol (PTZ)-Kindling response. (A) Female Gnao1+/G203R did not show heightened sensitivity to PTZ injection. (B) Male Gnao1+/G203R mice developed seizures earlier after repeated PTZ injections (Mantel-Cox Test; p<0.05). 3.4.3.2 R209H mutant mice are not hypersensitive to PTZ kindling Repeated application of a sub-threshold convulsive stimulus, leads to the generation of full-blown convulsions (Dhir, 2012). GNAO1 variants differ in their ability to cause epileptic seizures in patients. Children carrying the R209H mutant allele do not exhibit a seizure disorder. In accordance with the patients’ pattern, Gnao1+/R209H mice did not show increased susceptibility to kindling-induced seizures (Figure 3.13A & 3.13B). This contrasts with the increased kindling sensitivity in male G203R mutant mice (Figure 3.13) (Feng et al., 2019; Larrivee et al., 2019). 157! Figure 3.13 Gnao1+/R209H mice do not have an enhanced pentylenetetrazol (PTZ) kindling response. (A&B) Neither male nor female Gnao1+/R209H mice showed significant differences in sensitivity to PTZ injection compared to WT littermates (n.s.; Mantel-Cox test). 3.5 Discussion In this chapter, I describe three newly developed Gnao1 mutant mouse models (ΔT191F197, G203R, and R209H) and compare them with two previously published mouse models (G184S and KO). These data verify the genotype-phenotype correlation that I describe in chapter 2. Also, among the three different newly developed mouse models, only two mouse models (G203R and R209H) produced viable strains (Table 158! 3.3). Through the established behavioral pattern of the mouse models, we intend to explore mechanisms of Gnao1-associated movement disorders in the next chapter. Heterozygous male mice carrying the G203R mutation (GOF) in Gnao1 exhibit both a mild increase in seizure propensity and evidence of abnormal movements. This fits precisely with the variable seizure pattern of the children who carry this mutation as well as their severe choreo-athetotic movements (Arya et al., 2017; Dietel, 2016; Feng et al., 2018; Nakamura et al., 2013; Saitsu et al., 2016; Schirinzi et al., 2019; Schorling et al., 2017; Xiong et al., 2018). Heterozygous mice carrying the R209H mutation (NF) only develop hyperkinetic movements without loss of motor coordination on RotaRod or loss of capability of lifting heavy weights (Figure 3.8). This mimics patients with R209H mutations (Ananth et al., 2016; Blumkin et al., 2018; R. M. Dhamija, J, W.; Shah, B, B.; Goodkin, H, P., 2016; Kelly et al., 2019; Kulkarni et al., 2016; Marecos et al., 2018; Menke et al., 2016). In comparison, mice model with the LOF mutation ΔT191F197 developed spontaneous seizures at an early age and died prematurely before P16. This fits the clinical description of the patient carrying the same mutation (Nakamura et al., 2013) who died at 11 months (Figure 3.4). A summary of phenotypes of viable mutant mice is shown in Table 3.4. For comparison purposes, we have also tested the previously reported GOF mutant mouse model (G184S) and KO mouse model. The G184S mouse model exhibits a sex-dependent movement disorder while KO mice did not show any severe motor disability. 159! Table 3.4 Phenotypes of Gnao1 mutant mice Open Field Balance (RotaRod) Grip Strength Gait Analysis Seizure Susceptibility G203R Normal R209H G184S Hyperactivity Hypoactivity Reduced Normal Reduced KO Normal Normal Reduced Normal Reduced Normal ↓↓ stride length ↓stride length ↑ variability Increased Normal Increased ↓ stride length ↑↑ variability ↓ stride length Normal (Kehrl et al., 2014) (Kehrl et al., 2014) In mouse models of movement disorders, the mouse phenotype is usually not as striking or as easily observed as the clinical abnormalities in the patients (Oleas, Yokoi, DeAndrade, Pisani, & Li, 2013; Wilson & Hess, 2013), however they are often informative about mechanism and therapeutics. The male Gnao1+/G203R mutant mouse carrying patient-derived mutation very closely replicates the mild seizure phenotype of female Gnao1+/G184S mice (Kehrl et al., 2014). I now show that the female Gnao1+/G184S mice also exhibit gait and motor abnormalities. Both the GNAO1 G203R and the G184S mutations show a definite but modest GOF phenotype in biochemical measurements of cAMP regulation (Feng et al., 2017). In each case, the maximum percent inhibition of cAMP is not greatly increased but the potency of the a2A adrenergic agonist, used in those studies to reduce cAMP levels, was increased about 2-fold (Feng et al., 2017). This effectively doubles signaling through these two mutant G proteins at low 160! neurotransmitter concentrations (i.e. those generally produced during physiological signaling). This, however, does not prove that cAMP is the primary signal mechanism involved in pathogenesis of the disease. The heterotrimeric G protein, Go, is the defining subunit to many different effectors (Feng et al., 2018; Strittmatter et al., 1994; Wettschureck & Offermanns, 2005). We recently reviewed the mutations associated with genetic movement disorders and identified both cAMP regulation and control of neurotransmitter release as two mechanisms that seem highly likely to account for the pathophysiology of GNAO1 mutants (Chapter 1) (Feng et al., 2018). Since many Go signaling effectors (including cAMP and neurotransmitter release) can be mediated by the Gbg subunit released from the Go heterotrimer, other effectors could also be involved in the disease mechanisms. A recent hypothesis has also been raised that intracellular signaling by Gao may be involved (Solis & Katanaev, 2018). The observation that one of the most common movement disorder-associated alleles (R209H and other mutations in Arg209) does not markedly alter cAMP signaling in in vitro models, does suggest that the mechanism is more complex than a simple GOF vs LOF distinction at cAMP regulation. The R209H mutation was only tested for regulation of cAMP levels. It remains an unanswered question why a NF mutation still would lead to movement disorder in human patients and hyperactivity in our mouse models. This is a potential drawback of our in vitro assessment of cAMP only in an engineered HEK293T cell system. Since Gao regulates at least six different pathways (Jiang & Bajpayee, 2009), cAMP may not be the 161! affected downstream target of the R209H mutation. Therefore, the R209H animal model should be more valuable foe mechanistic studies since it will provide a more relevant physiological environment for studying the regulation of Gao in isolated neurons. Another interesting observation lies in the comparison between the KO mouse model and the LOF mouse model ΔT191F197. Although ΔT191F197 proved to be an epileptogenic and lethal mutation in both mouse and human, the actual KO mouse model did not develop any obvious seizure phenotype. Previously, homozygous Gnao1 KO mice were reported with a mild seizure phenotype (Jiang et al., 1998), however, heterozygous KO mice were relatively normal. One explanation could be the compensation effect of Gai protein, which takes over the mechanistic pathways that were once regulated by Gao protein. More likely, ΔT191F197 has some unknown mechanism, which would lead to the abnormal fetal development and infantile lethality. We also observed a striking sex difference in the phenotypes of our mouse models. Female Gnao1+/G184S mice and male Gnao1+/G203R mice showed much more prominent movement abnormalities than male G184S and female G203R mutants. However, the patterns of changes in the behavioral tests did not exactly overlap. G184S mutants showed significant changes in open field tests while G203R mutants did not. Conversely, G203R mutants showed a striking reduction in ability to walk/run at higher treadmill speeds while G184S mutants did not. For both mutant alleles, the seizure phenotype was also worse in the sex with more prominent movement disorder. For the NF mutant 162! line, Gnao1+/R209H mice did not show any sex difference in their hyperkinetic movemements in the open field arena (Figure 3.8), but they did have male dominated gait abnormalities, shown as decreased stride length (Figure 3.9). GNAO1 encephalopathies are slightly more prevalent (60:40) in female than male patients (Feng et al., 2018). It is not uncommon to have sex differences in epilepsy or movement disorder disease progression. One possible explanation is that estrogen prevents dopaminergic neuron depletion (Smith & Dahodwala, 2014). The Gi/o coupled estrogen receptor, GPR30, contributes to estrogen physiology and pathophysiology (Revankar, Cimino, Sklar, Arterburn, & Prossnitz, 2005). Also, PD is more common in male than female human patients (Wooten, Currie, Bovbjerg, Lee, & Patrie, 2004), therefore, the pro-dopaminergic properties of estrogen may exacerbate conditions mediated by hyper-dopaminergic symptoms like chorea in Hungtington’s disease (HD) (Smith & Dahodwala, 2014). Chorea/athetosis is the most prevalent movement pattern seen in GNAO1-associated movement disorders (Feng et al., 2018) so the female predominance correlates with that in HD. Clearly mechanisms of sex differences are complex including differences in synaptic patterns, neuronal densities and hormone secretion (Gillies, Murray, Dexter, & McArthur, 2004; Kompoliti, 1999; Smith & Dahodwala, 2014), but it is beyond the scope of this chapter to explain how the molecular differences contribute to the distinct behavioral patterns. A more detailed analysis on sex difference is provided in Chapter 5. 163! Since GNAO1 encephalopathy is often associated with developmental delay and cognitive impairment (Feng et al., 2018), it would be interesting to see whether the movement phenotype we have seen in female Gnao1+/G184S, male Gnao1+/G203R or Gnao1+/R209H mice is due to a neurodevelopmental malfunction or to ongoing active signaling alterations. Go-coupled GPCRs play an important role in hippocampal memory formation (Madalan, Yang, Ferris, Zhang, & Roman, 2012; Schutsky, Ouyang, & Thomas, 2011). Additional behavioral tests will be valuable to assess the learning and memory ability of the Gnao1 mutant mice. With the increasing recognition of GNAO1-associated neurological disorders, it is important to learn about the role of Go in the regulation of central nervous system. The novel Gnao1 G203R and R209H mutant mouse models reported here, and further models under development, should facilitate our understanding of GNAO1 mechanisms in the in vivo physiological background rather simply in in vitro cell studies. The animal models can also be used for preclinical drug testing and may permit a true allele-specific personalized medicine approach in drug repurposing for the associated movement disorders or epilepsy. 164! APPENDIX 165! APPENDIX SUPPLEMENTAL DATA Figure S3.1 RotaRod test was conducted with 5 training sessions and 1 test session over two consecutive days. (A) Female Gnao1+/G184S mice showed significantly motor abnormalities in test trial at day 2 (unpaired t-test; ***p<0.001). (B) Male Gnao1+/G184S mice did not show any significance in any training or test session. (C) Female Gnao1+/G203R mice did not exhibit any motor abnormalities in any RotaRod trial or test session. (D) Male Gnao1+/G203R mice showed significantly decreased capability in motor balance (unpaired t-test; *p<0.05). 166! A B F Gnao1+/+ F Gnao1+/G184S M Gnao1+/+ M Gnao1+/G184S 50 40 30 20 10 r e t e C n i i e m T % 15 10 15 10 0 F Gnao1+/+ F Gnao1+/G184S M Gnao1+/+ M Gnao1+/G184S 50 40 30 20 10 r e t e C n i i e m T % 0 F Gnao1+/+ F Gnao1+/G203R M Gnao1+/+ M Gnao1+/G203R 16 10 10 11 F Gnao1+/+ F Gnao1+/G203R M Gnao1+/+ M Gnao1+/G203R Figure S3.2 Time spent at the center in the Open Field Test. (A) No significant differences were observed between Gnao1+/G184S mice and their littermate controls. (B) No significant differences were observed between Gnao1+/G203R mice and their littermate controls. 167! Figure S3.3 RotaRod learning curve was collected in 10 consecutive tests with a 5-min break between each test. (A, C & E) Short-term learning curve comparison between Gnao1+/+ and Gnao1+/G203R in both sexes. (A & C) Both male and female Gnao1+/G203R mice showed reduced capability of keeping balance on RotaRod. (E) No significant difference in either sex between Gnao1+/+ and Gnao1+/G203R mice was observed comparing the rate of learning. (B, D & F) Short-term learning curve comparison between Gnao1+/+ and Gnao1+/G184S in both sexes. (B & D) Both male and female Gnao1+/G184S mice showed reduced capability of keeping balance on RotaRod. (F) No significant difference in either sexes between Gnao1+/+ and Gnao1+/G184S mice was observed comparing the rate of learning. 168! A Fore Limb F Gnao1+/+ vs Gnao1+/G184S 10 l e u a v r e t e m a r a P 1 Hind Limb F Gnao1+/+ vs Gnao1+/G184S B 100 10 1 l e u a v r e t e m a r a P 0.1 0.1 StrideLength MAX.dA.dT Swing Paw.Area.at.Peak.Stance.in.sq..cm Axis.Distance Overlap.Distance Stride.Frequency Stride Paw.Angle.Variability Fore Limb C M Gnao1+/+ vs Gnao1+/G184S 10 e u l a v r e t e m a r a P 1 Paw.Drag Propel Stance.Swing StrideLength MAX.dA.dT Swing Paw.Area.at.Peak.Stance.in.sq..cm Overlap.Distance Stance Stride.Frequency Stride Midline.Distance Paw.Angle.Variability Hind Limb 0.01 D M Gnao1+/+ vs Gnao1+/G184S 100 10 1 l e u a v r e t e m a r a P 0.1 MAX.dA.dT 0.1 Paw.Area.at.Peak.Stance.in.sq..cm Paw.Angle.Variability Paw.Drag MAX.dA.dT Tau...Propulsion 0.01 Swing Paw.Area.at.Peak.Stance.in.sq..cm Figure S3.4 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/G184S mice. All parameters that showed significance at belt speed 25 cm/s are plotted. (A&B) Female Gnao1+/G184S and their littermate controls showed parameters with significance detected by the FDR analysis. (C&D) Male Gnao1+/G184S and their littermates controls showed parameters with significance detected by the FDR analysis. FDR is calculated by a two-stage step-up method of Benjamini, Krieger and Yekutiel. Significant values are defined as q < 0.01. 169! A Hind Limb 1000 F Gnao1+/+ vs Gnao1+/G203R 100 10 l e u a v r e t e m a r a P 1 0.1 0.01 Stance.Swing StrideLength Swing Overlap.Distance Tau..Propulsion X.StanceStride X..Shared.Stance Midline.Distance X.SwingStride M Gnao1+/+ vs Gnao1+/G203R Hind Limb C 1000 100 l e u a v r e t e m a r a P 10 1 B Fore Limb M Gnao1+/+ vs Gnao1+/G203R 10 1 l e u a v r e t e m a r a P 0.1 0.1 0.01 Swing StrideLength Stride.Frequency Stride Absolute.PawAngle Overlap.Distance Stride.Length.CV X.Steps 0.01 Propel Stance.Swing StrideLength X.BrakeStance MAX.dA.dT Paw.Area.at.Peak.Stance.in.sq..cm Swing Brake X.BrakeStride Overlap.Distance X.StanceStride X..Shared.Stance Stride.Length.CV X.PropelStance X.Steps Swing.Duration.CV Midline.Distance Paw.Angle.Variability X.SwingStride Figure S3.5 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/G203R mice. All parameters that showed significance are plotted here. (A) Female Gnao1+/G203R and their littermate controls showed 9 parameters with significance only in hind limb data detected by the FDR analysis. (B&C) Male Gnao1+/G203R and their littermates controls exhibited 27 parameters with significance detected by the FDR analysis in fore and hind limb data combined. FDR is calculated by a two-stage step-up method of Benjamini, Krieger and Yekutiel. Significant values are defined as q < 0.01. 170! A Fore Limb F Gnao1+/+ vs Gnao1+/R209H F Gnao1+/+ vs Gnao1+/R209H Hind Limb B 1000 100 10 1 l e u a v r e t e m a r a P 0.1 0.01 C 10 1 l e u a v r e t e m a r a P 0.1 0.01 Overlap.Distance Fore Limb M Gnao1+/+ vs Gnao1+/R209H D Propel X.PropelStride X.BrakeStance X.PropelStance Swing Swing.Duration.CV Stride.Frequency Midline.Distance Hind Limb M Gnao1+/+ vs Gnao1+/R209H e u l a v r e t e m a r a P 10 1 0.1 0.01 Paw.Drag Overlap.Distance 1000 100 e u l a v r e t e m a r a P 10 1 Overlap.Distance 0.1 Paw.Drag MAX.dA.dT Paw.Area.at.Peak.Stance.in.sq..cm Overlap.Distance Figure S3.6 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/R209H mice. All parameters that showed significance are plotted here. (A&B) Female Gnao1+/R209H and their littermate controls showed 9 parameters with significance detected by the FDR analysis. (C&D) Male Gnao1+/R209H and their littermates controls exhibited fewer parameters with significance comparing to female detected by the FDR analysis in fore and hind limb data combined. FDR is calculated by a two-stage step-up method of Benjamini, Krieger and Yekutiel. Significant values are defined as q < 0.01. 171! M Gnao1+/+ vs Gnao1+/KO Fore Limb M A 100 l e u a v r e t e m a r a P 10 1 0.1 0.01 Swing X.BrakeStride StrideLength SLVar B 1000 100 l e u a v r e t e m a r a P 10 1 0.1 0.01 Hind Limb M M Gnao1+/+ vs Gnao1+/KO Swing X.SwingStride X.StanceStride Stance.Swing StrideLength Absolute.PawAngle Paw.Drag SLVar Paw.Area.Variability.at.Peak.Stan Swing.Duration.CV Midline.Distance Figure S3.7 False discovery rate (FDR) calculation probed of significantly different parameters from the DigiGait data in Gnao1+/- mice. All parameters that showed significance are plotted here. (A&B) Female Gnao1+/- and their littermate controls did not show any significant difference in any parameters given. Male Gnao1+/- and their littermates controls exhibited several parameters with significance comparing to female detected by the FDR analysis in fore and hind limb data combined. FDR is calculated by a two-stage step-up method of Benjamini, Krieger and Yekutiel. Significant values are defined as q < 0.01. 172! Table S3.1 Gait analysis parameters of male Gnao1 G203R mutant mice Feng et al Table S1 Gait analysis parameters Male Gnao1 G203R mutants Measured Parameter Fore Limb p X.SwingStride Swing Brake X.BrakeStride Propel X.PropelStride Stance X.StanceStride Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle Absolute.PawAngle Paw.Angle.Variability StanceWidth StepAngle SLVar SWVar StepAngleVar X.Steps Stride.Length.CV Stance.Width.CV Step.Angle.CV Swing.Duration.CV Paw.Area.at.Peak.Sta nce.in.sq..cm Paw.Area.Variability. at.Peak.Stan Hind.Limb.Shared.St ance.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositi oning.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag 0.000086 0.560859 0.653649 0.060086 0.003957 0.105634 0.020015 0.560859 0.001621 0.073312 0.073312 0.496204 <0.000001 0.002659 0.255886 0.000619 FDR Yes No No No No No No No Yes No No No Yes Yes No Yes 0.477312 0.50287 0.530415 0.19286 0.753361 0.566174 0.000053 0.000167 0.957513 0.205403 0.004977 0.111782 0.534568 0.656453 0.034845 0.926219 0.31531 0.000567 0.009576 0.065892 0.000511 0.813623 No No No No No No Yes Yes No No No No No No No No No Yes No No Yes No Hind Limb p <0.000001 0.000966 0.000399 <0.000001 0.006698 0.088909 0.413195 0.000966 0.010476 0.000004 0.000004 0.000155 <0.000001 0.024506 0.576762 0.7785 0.000239 0.718181 0.990996 0.158581 0.324024 0.422419 0.000013 0.001176 0.256205 0.276243 0.001453 0.004248 0.074908 0.286417 0.000002 0.927493 0.034845 0.001727 0.646552 0.455592 0.000567 0.009576 0.031149 0.000002 0.774606 0.013423 FDR Yes Yes Yes Yes Yes No No Yes No Yes Yes Yes Yes No No No Yes No No No No No Yes Yes No No Yes Yes No No Yes No No Yes No No Yes No No Yes No No Fore Limb M Gnao1+/G203R 0.08760833 37.85083333 0.06865 29.43125 0.076825 32.71791667 0.1454875 62.14916667 0.23310833 47.23583333 52.76416667 1.67083333 5.60791667 4.49708333 -1.50833333 5.14166667 8.31791667 4.51666667 86.1125 1.28591667 17.77083333 92.2875 21.81875 23.578375 83.45 96.21666667 29.98779167 M Gnao1+/+ 0.09211667 38.07777778 0.06939444 28.40222222 0.08235556 33.52 0.15173889 61.92222222 0.24383333 45.82444444 54.17555556 1.65222222 6.09777778 4.30055556 -2.13944444 3.90722222 8.12333333 4.77777778 92 1.24272222 18.37222222 85.73888889 24.00833333 20.82244444 82.95 81.05555556 27.594 SD 0.0111967 3.54836583 0.016458 4.89821488 0.01995907 4.68472378 0.02844761 3.54836583 0.03618381 7.20043157 7.20043157 0.25378818 0.77998392 0.63908954 4.87883924 3.61305873 2.56029197 4.11368551 97.32018227 0.32258874 19.78855562 114.0590048 4.7819712 6.52396481 94.95293012 111.6869989 7.89706951 0.30805556 0.04937709 0.02894444 0.01608279 0.2975 0.07788791 0.02995833 0.01688003 1 1 1 0 0 0 11.30555556 1.01388889 16.896 -5.16177778 11.54350393 0.04501259 3.37636885 1.45335565 11.8375 1.02416667 16.8605 -5.31579167 12.53391424 0.05217581 4.22700441 1.6247547 1.4025 0.4440782 1.55470833 0.44461341 0.47255556 0.899 -2.22233333 0.01011111 1 0.209826 0.30885281 0.34125483 0.83342464 0 0.53175 0.95725 -2.37679167 -0.00895833 0.24509639 0.328734 0.51249994 0.80948622 0 1 1 1 1 n 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 SD 0.01176783 4.2332119 0.01707632 5.96865439 0.01889352 5.25089869 0.02613979 4.2332119 0.03278618 8.50521667 8.50521667 0.29341328 0.74607749 0.67417378 6.12588422 3.64193574 2.92469857 3.82128663 93.39500767 0.34549218 19.10096449 116.870725 5.87860657 7.92635878 95.25775208 127.9222556 9.08822921 0 0 0 n 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 M Gnao1+/+ 0.09002222 36.68666667 0.03542222 14.345 0.12215 48.96611111 0.15759444 63.31333333 0.24762222 22.61888889 77.38111111 1.75555556 6.17777778 4.25111111 0.39222222 16.34777778 4.79555556 9.53888889 46.90555556 0.89744444 8.52777778 83.06111111 23.67777778 14.70616667 128.2111111 99.85555556 21.31672222 SD 0.01187409 3.98361729 0.0096027 3.47436381 0.0268225 4.21758972 0.03072707 3.98361729 0.03807229 5.02721964 5.02721964 0.3053018 0.72670955 0.64785757 17.03757417 4.65733584 1.82575682 8.65887596 62.27966276 0.27275979 12.16526356 110.7694855 4.70635557 4.90287218 154.5150275 121.4064747 6.46662759 0.64966667 0.12237364 0.05305556 0.033092 22 118.0833333 13.59444444 1.01388889 46.49827778 -8.88127778 178.9333333 1.4025 0.47255556 0.63983333 1.76005556 0.02372222 218.8277778 26.92406039 67.80483618 13.27605364 0.04501259 9.49935633 1.91057189 101.5146521 0.4440782 0.209826 0.24794907 0.28026224 1.34792041 112.0513066 n 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 180 Hind Limb M Gnao1+/G203R 0.08343333 35.24083333 0.0396125 16.53625 0.115625 48.225 0.15522917 64.75916667 0.23864167 25.36166667 74.63875 1.88625 5.73458333 4.39541667 1.335 16.2 5.57333333 9.85416667 46.975 0.95095833 9.69583333 92.04583333 21.32291667 16.811125 111.1208333 113.7708333 24.35395833 SD 0.01239566 4.70652526 0.01337189 4.86819351 0.0221926 4.54530426 0.02815843 4.70652526 0.03329752 6.5864662 6.58555209 0.375531 0.71873598 0.64873536 17.17832949 5.77465245 2.32933748 8.99741847 62.43865106 0.45003245 11.87042946 115.4575489 5.88566941 7.52811028 150.8803385 135.1570046 11.41084435 0.61033333 0.14986149 0.05908333 0.03507305 24.97916667 152.3458333 13.475 1.02416667 43.518625 -8.980875 186.7458333 1.55470833 0.53175 0.7065 1.57204167 -0.01433333 190.6458333 29.29998471 74.75690858 13.32500873 0.05217581 9.64248797 2.39585688 109.3986127 0.44461341 0.24509639 0.35345432 0.46433242 1.34621132 117.356213 n 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 240 173! Table S3.2 Gait analysis parameters of female Gnao1 G203R mutant mice Gait analysis parameters Female Gnao1 G203R mutants Feng et al Table S2 n 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 F Gnao1+/+ 0.08701905 37.98428571 0.03703333 15.94428571 0.10746667 46.06619048 0.14451429 62.01571429 0.23155238 25.63047619 74.36952381 1.66857143 5.80761905 4.54809524 0.40666667 17.21047619 5.24428571 8.97619048 46.68095238 0.82042857 7.22380952 93.55714286 23.34761905 14.45657143 127.6666667 96.90952381 20.35714286 SD 0.0103341 4.4849659 0.0117622 4.3353439 0.0239663 4.7852095 0.029293 4.4849659 0.0346545 6.3902157 6.3902157 0.3092867 0.7083501 0.6820467 18.071952 5.3985099 2.5795276 8.1636612 62.270502 0.3400398 9.302076 122.72527 5.9556269 6.7501268 155.40339 123.48735 10.045441 0.624 0.138264 0.06057143 0.0355498 19.06190476 116.7095238 13.52380952 1.0192381 43.48771429 -9.69980952 164.8380952 1.42585714 0.44066667 0.62028571 1.8667619 -0.004 212.1142857 23.988055 70.404616 13.510298 0.0513119 9.941629 2.6687581 101.35854 0.3488621 0.1970416 0.3216493 0.3241286 1.2942361 132.61357 n 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 Hind Limb F Gnao1+/G203R 0.08259735 36.3 0.03834513 16.8199115 0.10812832 46.88185841 0.14648673 63.70044248 0.22903982 26.23938053 73.76061947 1.79070796 5.5659292 4.59734513 1.02699115 15.92699115 5.44778761 8.96902655 53.61946903 0.86349558 8.37610619 89.84955752 24.08628319 15.80809735 123.6283186 105.039823 22.57513274 SD 0.011866 3.9776487 0.0114719 4.766196 0.0226036 4.0616338 0.0268555 3.9787156 0.0339969 6.5617814 6.5617814 0.3146108 0.6710776 0.6955682 16.942806 5.7724885 3.1063833 8.1553754 67.358321 0.299672 10.810084 112.67252 5.4387774 6.1679082 157.26751 125.67239 8.4641854 0.60482301 0.1192093 0.05575221 0.029599 21.98230088 144.300885 13.12389381 1.01123894 42.92628319 -8.8219469 206.1415929 1.30376106 0.47349558 0.69393805 1.61292035 -0.00207965 215.9734513 25.48742 66.869808 12.926894 0.0292767 8.6321179 2.4798691 106.91923 0.3572146 0.2414856 0.3180957 0.3222468 1.2662788 114.3211 n 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 226 11.6952381 12.247545 1.0192381 0.0513119 16.9396667 4.1600988 -5.1652381 1.8030427 12.27876106 1.01123894 16.29477876 -4.63566372 12.170536 0.0292767 3.8423131 1.6700964 1.42585714 0.3488621 1.30376106 0.3572146 1 1 1 0 0 0 1 1 1 1 0 0 0 0 Fore Limb F Gnao1+/G203R 0.08465044 37.58362832 0.07067699 31.14867257 0.07115487 31.26681416 0.14184071 62.41637168 0.22645575 49.8840708 50.1159292 1.68141593 5.50530973 4.64778761 -1.51504425 4.13274336 7.53495575 4.28318584 83.71238938 1.13561947 15.69026549 88.42920354 24.34070796 21.13402655 82.69469027 100.6283186 27.70469027 SD 0.011645 3.1433241 0.0163127 5.0584954 0.0169689 4.8963937 0.0252212 3.1433241 0.0342293 7.5687729 7.5687729 0.2274687 0.6725528 0.7174148 4.944406 3.0979403 2.560203 3.588795 92.843532 0.3330556 16.365318 114.15516 5.3805268 7.3727667 91.571949 129.68952 8.038455 0.30084071 0.0747244 0.03048673 0.0186722 SD 31.35 F Gnao1+/+ 0.0862619 0.0112876 38.2376191 3.4885409 0.06979524 0.019761 30.4114286 5.6522649 0.07122857 0.0153668 4.8383758 0.14104286 0.0265394 61.762381 3.4885409 0.22737619 0.0347068 8.051334 49.1166667 50.8833333 8.051334 1.63809524 0.2335009 5.70190476 0.7409397 4.63333333 0.7236493 -2.0414286 5.1145489 4.27380952 3.4631111 7.08 2.9170313 4.65238095 3.9041285 87.4904762 95.888554 1.11104762 0.3637969 14.8666667 16.15035 94.5571429 123.21106 23.7642857 5.9102466 20.0334286 7.9432229 71.4761905 83.50877 92.3857143 125.23065 25.4634286 9.9336491 0.31547619 0.0784235 0.03247619 0.0177047 0.44066667 0.1970416 0.85080952 0.36314 -1.9551905 0.4090709 -0.0137143 0.794259 n 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 210 Fore Limb Hind Limb Measured Parameter X.SwingStride Swing Brake X.BrakeStride Propel X.PropelStride Stance X.StanceStride Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle Absolute.PawAngle p 0.143574 0.040102 0.610655 0.151434 0.962205 0.858596 0.747729 0.040102 0.780626 0.30554 0.30554 0.050429 0.003872 0.83429 0.275233 0.653751 FDR No No No No No No No No No No No No No No No No p 0.000042 0.000039 0.239211 0.045932 0.766869 0.055104 0.463647 0.000039 0.445327 0.327424 0.327424 0.000053 0.000285 0.456267 0.711622 0.017128 FDR Yes Yes No No No No No Yes No No No Yes Yes No No No SLVar SWVar StanceWidth StepAngle Paw.Angle.Variability 0.083677 0.304131 0.676222 0.46197 0.597496 0.590118 0.287034 0.134218 0.183108 0.500567 0.009699 Stride.Length.CV Stance.Width.CV Step.Angle.CV StepAngleVar Swing.Duration.CV Paw.Area.at.Peak.Sta X.Steps nce.in.sq..cm Paw.Area.Variability. at.Peak.Stan Hind.Limb.Shared.St ance.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositi oning.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag 0.046636 0.25505 0.618239 0.044325 0.093152 0.001558 0.000346 0.122248 0.125326 0.844898 0.882318 No No No No No No No No No No No No No No No No No No No No No No 0.459013 0.992695 0.265709 0.160659 0.235126 0.742411 0.176573 0.029457 0.787718 0.496458 0.012817 0.120876 0.123726 0.219455 0.000033 0.752284 0.044325 0.528498 0.000411 0.000043 0.000346 0.122248 0.01669 <0.000001 0.987517 0.744503 No No No No No No No No No No No No No No Yes No No No Yes Yes Yes No No Yes No No 0.47349558 0.90159292 -1.96256637 -0.00261062 1 0.2414856 0.3272514 0.3777178 0.7706896 0 174! Table S3.3 Gait analysis parameters of male Gnao1 G184S mutant mice Gait analysis parameters Male Gnao1 G184S mutants Feng et al Table S3 Fore Limb M Gnao1+/G184S 0.08857143 37.6702381 0.05385714 22.75119048 0.09477381 39.57857143 0.14857143 62.3297619 0.23721429 36.49047619 63.50952381 1.67142857 5.93333333 4.44761905 0.91547619 4.55119048 8.06666667 3.79761905 46.78571429 1.32916667 14.6547619 39.98809524 22.25595238 22.82166667 35.82142857 40.55952381 26.30416667 SD 0.011722 3.3344055 0.0134703 4.3988359 0.023563 4.5858365 0.0298184 3.3344055 0.0389603 6.6003216 6.6003216 0.2341719 0.693226 0.7554582 5.4220321 3.046922 3.1549286 3.0528666 50.288342 0.4364264 15.444026 52.336606 5.1316958 8.3634236 40.621426 53.357704 9.8202879 0.26928571 0.0702111 0.02559524 0.0187148 n 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 n 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 M Gnao1+/+ 0.09 36.51428571 0.02879592 11.57857143 0.13060204 51.91632653 0.15944898 63.48571429 0.24947959 18.21020408 81.78979592 1.79183673 6.26122449 4.22346939 -0.05204082 17.08877551 4.2877551 7.12244898 25.30612245 0.91683673 6.80612245 42.21428571 21.0255102 14.93540816 57.7244898 48.95918367 21.32826531 SD 0.0127724 5.1117008 0.0075135 2.4412372 0.029081 4.6134205 0.0336318 5.1117008 0.0391545 3.3839064 3.3839064 0.3701974 0.9164738 0.6697901 17.660751 4.1070692 1.9761081 6.2363506 32.434997 0.309591 7.6934276 51.158184 4.4032395 5.4905719 69.576003 59.29655 7.197867 0.70244898 0.1316122 0.04530612 0.030127 14.24489796 74.91836735 9.25510204 1.01755102 51.415 -8.24183673 103.755102 1.40122449 0.43704082 0.61469388 1.40795918 0.01989796 107.2653061 17.100709 45.56058 9.6744016 0.0511529 10.373366 2.1764496 45.984547 0.3999259 0.1825829 0.248655 0.2798857 1.3185305 45.749779 n 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 98 Hind Limb M Gnao1+/G184S 0.08357143 35.3 0.02907143 12.08333333 0.12857143 52.62738095 0.15763095 64.7 0.24122619 18.70238095 81.29880952 1.89642857 6.01071429 4.40238095 0.675 16.66309524 5.24047619 7.78571429 22.44047619 0.9572619 7.66666667 36.42857143 21.95238095 16.00988095 59.47619048 58.16666667 23.49071429 SD 0.0102043 5.2765451 0.0081415 2.8270992 0.0333879 5.3388179 0.0377269 5.2765451 0.0423552 4.2032969 4.2045728 0.4203975 0.634551 0.7735034 17.87932 6.2551719 2.7850772 7.0198882 30.489718 0.4562322 9.442704 44.194064 5.1569459 7.6712154 70.164948 63.81578 11.914931 0.60595238 0.1286643 0.05357143 0.0337833 15.35714286 87.82142857 9.55952381 1.01452381 44.32047619 -9.5402381 72.35714286 1.61261905 0.51178571 0.6727381 1.51952381 0.00738095 72.89285714 17.924637 42.745862 9.5328869 0.0503579 10.399373 2.7031744 49.842316 0.5000846 0.1905993 0.3411474 0.4244029 1.4239324 52.00487 n 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 84 Fore Limb Hind Limb Measured Parameter X.SwingStride X.BrakeStride X.PropelStride X.StanceStride Swing Brake Propel Stance Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle Absolute.PawAngle p 0.043466 0.726843 0.826349 0.180502 0.148248 0.318489 0.27969 0.726843 0.140428 0.197014 0.197014 0.810208 0.047635 0.153098 0.223604 0.062068 FDR No No No No No No No No No No No No No No No No p 0.000278 0.117241 0.812719 0.197797 0.66152 0.336379 0.731504 0.117241 0.173925 0.382853 0.384084 0.075995 0.036347 0.096182 0.783411 0.583264 FDR Yes No No No No No No No No No No No No No No No SLVar SWVar StanceWidth StepAngle Paw.Angle.Variability 0.000004 0.931208 0.781449 0.018345 0.268206 0.51712 0.201763 0.009988 0.245913 0.386308 0.495053 Stride.Length.CV Stance.Width.CV Step.Angle.CV StepAngleVar Swing.Duration.CV Paw.Area.at.Peak.Sta X.Steps nce.in.sq..cm Paw.Area.Variability. at.Peak.Stan Hind.Limb.Shared.St ance.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositi oning.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag <0.000001 0.002006 0.396136 0.688994 <0.000001 0.891498 0.001809 0.00764 0.055735 0.998012 0.836286 Yes No No No No No No No No No No Yes No No No Yes No No No No No No 0.007857 0.50059 0.542101 0.480115 0.49907 0.419338 0.192532 0.274044 0.866251 0.314724 0.13408 0.000001 0.082786 0.669308 0.051594 0.831523 0.688994 0.000008 0.000429 0.000017 0.001809 0.00764 0.187315 0.035382 0.951005 0.000004 No No No No No No No No No No No Yes No No No No No Yes Yes Yes Yes No No No No Yes M Gnao1+/+ 0.09253061 37.83979592 0.05344898 21.91020408 0.09986735 40.25714286 0.15326531 62.16020408 0.24587755 35.25306122 64.74693878 1.66326531 6.17959184 4.29387755 -0.00102041 3.73163265 6.31938776 3.83673469 48.89795918 1.18122449 12.31632653 45.3877551 21.35714286 19.70622449 29.3877551 33.97959184 25.39285714 SD 0.0141627 3.1936386 0.0115937 4.0355505 0.0236179 4.5416969 0.028498 3.1936386 0.0396769 6.2745798 6.2745798 0.2230934 0.9320147 0.6895115 4.7029821 2.8371467 1.7027969 3.0348283 51.835284 0.4015749 12.961161 58.864099 4.3328046 7.7656016 33.936154 48.807634 8.1613281 0.32091837 0.0605148 0.01877551 0.0098719 0.43704082 0.81897959 -2.76357143 0.03887755 1 0.1825829 0.3422138 0.3019703 0.8386733 0 1 1 1 0 0 0 8.70408163 1.01755102 19.58153061 -4.68867347 9.4859736 0.0511529 3.7431478 0.9273804 7.57142857 1.01452381 16.61440476 -4.71535714 8.2932484 0.0503579 3.9087463 1.6546185 1.40122449 0.3999259 1.61261905 0.5000846 1 1 1 0 0 0 0.51178571 0.91821429 -2.76369048 0.01333333 1 0.1905993 0.3516473 0.3417171 0.8200578 0 175! Table S3.4 Gait analysis parameters of female Gnao1 G184S mutant mice Gait analysis parameters Female Gnao1 G184S mutants Feng et al Table S4 Fore Limb Hind Limb Measured Parameter X.SwingStride X.BrakeStride X.PropelStride X.StanceStride Swing Brake Propel Stance Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle Absolute.PawAngle p 0.000002 0.352054 0.018172 0.605581 0.073513 0.303687 0.004809 0.352054 0.000183 0.490318 0.489918 0.335195 0.0007 0.000035 0.167612 0.167612 FDR Yes No No No No No No No Yes No No No Yes Yes No No p 0.00019 0.571929 0.041827 0.827244 0.003264 0.75231 0.001181 0.571929 0.000242 0.875453 0.875453 0.536831 0.000352 0.000096 0.464223 0.464223 FDR Yes No No No Yes No Yes No Yes No No No Yes Yes No No F Gnao1+/+ 0.08906667 38.21333 0.062156 26.46 0.083178 35.33444 0.145322 61.78667 0.234389 42.76444 57.23556 1.64 5.707778 4.468889 4.784444 4.784444 6.265556 1.684444 64.84444 1.152444 0.349556 13.42622 18.32778 20.47044 21.188 21.42844 25.59067 SD 0.012567 3.504418 0.017608 5.698922 0.019603 5.498685 0.026527 3.504418 0.035201 8.767346 8.767346 0.240692 0.807719 0.682797 2.854089 2.854089 2.252647 0.330259 7.514911 0.381574 0.098164 4.439098 5.488579 6.723017 6.590364 8.660338 8.521292 0.398111 0.152248 0.031556 0.019309 1.009333 1.012889 25.22511 6.588111 0.075179 0.039413 9.069926 2.658642 1.749778 0.686131 0.443222 0.794889 2.572 0.838222 1 0.221435 0.295526 0.38345 0.176753 0 n 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 SLVar SWVar StanceWidth StepAngle Paw.Angle.Variability 0.000023 0.022922 0.998777 0.164572 0.200099 0.882892 0.00456 0.568182 0.007756 0.734363 0.412289 Stride.Length.CV Stance.Width.CV Step.Angle.CV StepAngleVar Swing.Duration.CV Paw.Area.at.Peak.Sta X.Steps nce.in.sq..cm Paw.Area.Variability.a t.Peak.Stan Hind.Limb.Shared.Sta nce.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositio ning.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag 0.000002 0.291841 0.167306 0.25904 <0.000001 0.014025 0.001549 0.979063 0.203884 0.039565 0.000293 0.055735 0.998012 0.836286 Yes No No No No No No No No No No Yes No No No Yes No Yes No No No Yes No No No 0.003025 0.068763 0.466437 0.379246 0.091884 0.454172 0.006296 0.504883 0.02606 0.682633 0.494238 0.000015 0.294115 0.112442 0.28745 0.357268 0.25904 0.000014 0.119243 0.253295 0.001549 0.979063 0.38151 0.000029 0.015682 0.00008 Yes No No No No No No No No No No Yes No No No No No Yes No No Yes No No Yes No Yes Hind Limb F Gnao1+/G184S 0.082333 37.79697 0.026402 12.075 0.110394 50.11818 0.136818 62.20303 0.219182 19.34394 80.65606 1.667424 5.420455 4.823485 16.82121 16.82121 5.02197 2.410606 54.86515 0.757652 0.259242 12.06667 19.98864 14.31485 10.9447 22.30394 18.80689 SD 0.012478 3.394359 0.00874 3.589752 0.023384 3.746567 0.026794 3.394359 0.03649 5.293627 5.293627 0.242599 0.691799 0.795393 5.22535 5.22535 2.460662 0.412918 8.435848 0.273303 0.129024 3.238623 4.73676 5.874303 5.480517 6.351191 6.401079 0.660227 0.230086 0.059773 0.045725 0.051152 36.32273 1.019545 1.018333 51.24644 10.95538 0.138298 1.463712 0.442424 0.579848 1.190758 1.220076 5.45697 0.019511 8.859942 0.061129 0.032011 18.95255 5.069048 0.06125 0.629242 0.222697 0.279277 0.564684 0.219397 2.551102 n 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 n 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 132 F Gnao1+/+ 0.088633 37.53889 0.029 12.18 0.119811 50.28222 0.148811 62.46111 0.237511 19.45556 80.54444 1.687778 5.781111 4.415556 16.32778 16.32778 4.153333 2.566667 56.06222 0.793778 0.220667 12.55133 18.08889 13.79056 8.748444 22.82822 18.19433 SD 0.011621 3.24618 0.01003 3.40298 0.022828 3.871491 0.026549 3.24618 0.035089 5.068056 5.068056 0.237888 0.775306 0.680369 4.441025 4.441025 1.479538 0.475299 8.526091 0.335337 0.097547 3.478382 5.451305 5.541985 4.294501 6.986445 6.749598 0.818667 0.301436 0.067 0.056299 0.057378 37.59111 1.009556 1.012889 64.31333 12.05456 0.150978 1.749778 0.443222 0.546444 1.539222 1.299111 7.0439 0.021 8.462962 0.047095 0.039413 24.77673 5.245177 0.103392 0.686131 0.221435 0.277777 0.641965 0.261708 3.321218 n 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 90 Fore Limb F Gnao1+/G184S 0.080841 37.76212 0.056508 26.05303 0.078295 36.1803 0.134818 62.23788 0.215674 41.91894 58.08182 1.672727 5.342424 4.908333 4.209848 4.209848 7.730303 1.536364 64.84697 1.092273 0.375 13.54348 20.31818 20.97265 25.38455 22.01364 24.72242 SD 0.012049 3.562839 0.017191 5.79599 0.02004 6.320866 0.027276 3.562839 0.03649 9.074334 9.073969 0.252647 0.755956 0.810338 3.153645 3.153645 2.613352 0.333114 9.138359 0.261539 0.104679 3.867871 4.78319 6.21845 8.830206 9.055266 7.146609 0.311288 0.11312 0.029091 0.015356 0.988788 1.018333 19.5603 5.764015 0.077291 0.032011 6.647057 2.269378 1.463712 0.629242 0.442424 0.845379 2.450985 0.752045 1 0.222697 0.285901 0.455068 0.167454 0 176! Table S3.5 Gait analysis parameters of male Gnao1 R209H mutant mice N 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 M WT Mean 0.08889286 36.2732143 0.03811607 15.1651786 0.12082143 48.5633929 0.15891964 63.7267857 0.24777679 23.6625 76.3375 1.80178571 6.23035714 4.21964286 1.14821429 17.0785714 4.3625 7.05357143 31.9732143 0.83419643 5.17857143 46.75 24.3080357 13.5108036 65.6696429 52.7767857 19.8541964 SD 0.009719 4.576024 0.011977 3.560158 0.021481 3.628342 0.029695 4.576024 0.033051 4.618239 4.618239 0.371018 0.794104 0.559736 17.7401 4.6622 1.635481 6.172942 43.08831 0.271292 5.429672 60.6561 4.825909 4.479219 74.19895 64.92819 5.777298 0.57464286 0.063401 0.044375 0.019162 16.2678571 82.4196429 10.7142857 1.00232143 40.355625 -7.7479464 90.5625 1.03803571 0.41776786 0.58044643 1.52928571 0.01223214 108.973214 19.84035 46.81026 10.60293 0.026913 4.795198 1.135325 46.16808 0.333044 0.151887 0.24559 0.19725 1.234583 56.80795 N 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 Hind Limb M R209H Mean 0.08860417 35.76979167 0.03653125 14.56770833 0.12441667 49.66354167 0.1609375 64.23020833 0.2495625 22.54375 77.45729167 1.83645833 6.13854167 4.20104167 0.85520833 15.265625 4.95104167 7.17708333 28.34375 0.89395833 6.15625 48.04166667 24.65104167 14.95729167 76.25 54.9375 21.98125 SD 0.012193 4.245595 0.012519 4.416536 0.023144 4.175722 0.028273 4.245595 0.033977 5.984204 5.983552 0.379784 0.76066 0.578291 16.02805 4.705849 2.276886 6.412151 38.24184 0.326185 6.875323 58.69079 5.644336 6.409749 82.21333 64.68544 8.948453 0.53677083 0.068032 0.04375 0.019317 17.125 88.27083333 10.6875 0.9975 37.9240625 -7.51552083 104.5729167 1.251875 0.4609375 0.65375 1.506875 -0.01677083 96.90625 19.99961 45.32908 10.76477 0.044745 5.440301 1.551981 64.28223 0.401419 0.193277 0.31857 0.277653 1.264412 61.0565 N 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 96 Fore Limb Hind Limb P value 0.742713 0.186611 0.07883 0.046618 0.476735 0.224488 0.459252 0.186611 0.606739 0.144106 0.143904 0.254797 0.514326 0.599868 0.409773 0.056783 0.708628 0.384371 0.543298 0.435215 0.920715 0.24211 0.8498 0.286712 0.642162 0.453618 0.065786 FDR No No No No No No No No No No No No No No No No No No No No No No No No No No No P value 0.849568 0.414505 0.352604 0.281478 0.246967 0.043317 0.617997 0.414505 0.701773 0.130071 0.129692 0.507041 0.397668 0.814209 0.901338 0.005873 0.031802 0.887767 0.524421 0.15068 0.253516 0.876653 0.63707 0.057982 0.33056 0.810819 0.040295 FDR No No No No No No No No No No No No No No No No No No No No No No No No No No No 0.269777 0.906173 0.904653 0.339967 0.337523 0.043386 0.00004 0.072965 0.131163 0.193089 0.734326 No No No No No No Yes No No No No 0.000048 0.815503 0.757282 0.36289 0.985628 0.339967 0.000739 0.215019 0.069796 0.00004 0.072965 0.062689 0.49868 0.867511 0.141646 Yes No No No No No Yes No No Yes No No No No No Measured Parameters X.SwingStride X.BrakeStride X.PropelStride X.StanceStride Swing Brake Propel Stance Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle StanceWidth StepAngle SLVar SWVar StepAngleVar X.Steps Stride.Length.CV Stance.Width.CV Step.Angle.CV Absolute.PawAngle Paw.Angle.Variability Swing.Duration.CV Paw.Area.at.Peak.Stanc e.in.sq..cm Paw.Area.Variability.at.P eak.Stan Hind.Limb.Shared.Stan ce.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositioni ng.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag M WT Mean 0.09575 39.058929 0.0748482 30.158036 0.0765714 30.778571 0.1515089 60.941071 0.2472232 49.605357 50.395536 1.5875 6.2089286 4.2294643 -1.589286 3.075 6.4607143 3.4107143 48.8125 1.0917857 12.625 45.919643 24.424107 17.998393 31.107143 49.205357 24.166161 SD 0.010045 3.751313 0.018216 5.195713 0.02064 5.958209 0.02807 3.751313 0.03416 8.646812 8.64638 0.262395 0.784024 0.586429 3.516355 2.31787 2.427261 2.559397 54.21986 0.27152 12.67695 60.59437 4.986563 5.348469 36.02711 63.63878 6.335163 0.2744643 0.031902 0.0205357 0.010209 1 1 1 9.5178571 1.0023214 14.877321 -3.363571 9.432532 0.026913 2.009157 0.813505 0 0 0 0.4177679 0.7350893 -1.933214 0.0076786 1 0.151887 0.247932 0.331655 0.789782 0 1.0380357 0.333044 1.251875 0.401419 N 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 112 Fore Limb M R209H Mean 0.09525 38.39479167 0.07973958 31.809375 0.074625 29.796875 0.15432292 61.60520833 0.24967708 51.459375 48.540625 1.628125 6.1375 4.18645833 -1.11875 3.77708333 6.57083333 3.11458333 53.5625 1.13145833 12.44791667 56.60416667 24.5625 19.09291667 33.63541667 42.73958333 26.689375 SD 0.011893 3.422753 0.021724 6.68853 0.018382 5.594329 0.026341 3.422753 0.034296 9.584303 9.584303 0.247813 0.788636 0.590917 4.682854 2.962662 1.67997 2.298145 58.20288 0.449982 12.89053 70.76982 5.537076 9.18012 42.3313 59.83877 12.71546 0.2796875 0.036171 0.02072917 0.013396 1 1 1 9.67708333 0.9975 15.17510417 -3.645625 9.675248 0.044745 2.457478 1.177005 0 0 0 0.4609375 0.805625 -1.9984375 -0.02947917 1 0.193277 0.413488 0.388778 0.781835 0 177! Table S3.6 Gait analysis parameters of female Gnao1 R209H mutant mice Fore Limb Hind Limb Brake Stride Swing Propel Stance X.BrakeStride X.SwingStride X.PropelStride X.StanceStride Stride.Frequency Measured Parameters X.BrakeStance X.PropelStance Stance.Swing StrideLength P value 0.000862 0.553568 0.257048 0.239615 0.062971 0.457782 0.052934 0.553568 0.009436 0.314746 0.314746 0.554245 0.920542 0.014077 0.094114 0.008759 0.273075 0.48535 0.829913 0.178551 0.620186 0.52429 0.134323 0.237484 0.865848 0.162584 0.294853 Paw.Area.at.Peak.Stance.in.sq..cm 0.00641 Paw.Area.Variability.at.Peak.Stan 0.166619 Hind.Limb.Shared.Stance.Time Stride.Length.CV Stance.Width.CV Step.Angle.CV Absolute.PawAngle Paw.Angle.Variability StanceWidth StepAngle Swing.Duration.CV StepAngleVar SLVar SWVar PawAngle X.Steps FDR No No No No No No No No No No No No No No No No No No No No No No No No No No No No No X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag PawPlacementPositioning.PPP. 0.369433 0.140168 0.013856 0.020047 0.000061 0.110841 0.291223 0.495964 0.755914 No No No No Yes No No No No P value 0.001686 0.784138 0.198976 0.003087 0.000237 0.000572 0.019444 0.784138 0.003393 0.001422 0.001422 0.625985 0.777101 0.002251 0.886736 0.141453 0.075906 0.502257 0.607524 0.006484 0.655617 0.895657 0.121155 0.006673 0.556665 0.685476 0.002003 0.66547 0.005094 0.537429 0.630463 0.629805 0.140168 0.798156 0.170644 0.758007 0.000061 0.110841 0.016224 0.002418 0.937074 0.000004 F WT Mean 0.088759 38.269277 0.0732651 31.40241 0.0713374 30.322289 0.1446145 61.730723 0.2334036 50.877108 49.122892 1.6361446 5.836747 4.5030121 -1.226506 3.673494 6.5048193 3.5240964 62.024096 1.0621687 14.644578 61.03012 27.427711 18.675843 50.23494 51.463855 25.371325 0.2454217 0.0223494 1 1 1 1 9.3192771 1.0078313 12.997831 -3.557771 0.9884337 0.4084337 0.7913855 -1.795783 0.003253 SD 0.011161 3.355496 0.016007 4.642243 0.017694 4.635664 0.026021 3.355496 0.033972 6.954237 6.954237 0.235657 0.776261 0.657122 4.09998 2.178745 1.980413 2.863991 67.35433 0.280301 15.3228 78.19598 6.803579 6.063927 55.10375 72.44272 7.757021 0.046964 0.0077 0 0 0 0 9.282136 0.026056 2.686344 0.714287 0.392521 0.147262 0.282622 0.305625 0.750176 N 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 Fore Limb F R209H Mean 0.09418182 38.53295455 0.07582955 30.67613636 0.07560227 30.7875 0.15142045 61.46704545 0.24563636 49.93863636 50.06136364 1.61818182 5.82613636 4.28636364 -0.36022727 2.94204545 6.20113636 3.79545455 60.11363636 1.11465909 15.68181818 54.72727273 26.14772727 19.66806818 51.48863636 65.53409091 24.28715909 0.2625 0.02386364 1 1 1 1 8.26136364 1.01545455 13.86136364 -3.82125 1.23068182 0.44352273 0.83193182 -1.82375 -0.02784091 SD 0.013945 3.399754 0.019054 4.729754 0.016582 4.943423 0.027503 3.399754 0.038132 7.271317 7.271317 0.218955 0.859229 0.678264 3.519944 1.940655 2.301149 3.093037 67.41746 0.321265 16.8117 68.41115 5.759629 6.872459 58.28448 82.82733 7.973726 0.047398 0.009276 0 0 0 0 8.197973 0.055974 2.557019 1.069478 0.543744 0.197449 0.305505 0.321102 0.771926 N 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 SD F WT Mean 0.0102 0.0841024 4.11783 36.083735 0.013145 0.0411506 4.881231 17.425904 0.021791 0.1098855 3.90007 46.489157 0.028399 0.1510843 4.11783 63.916265 0.03372 0.2352108 6.667062 27.063855 6.667062 72.936145 0.324599 1.8054217 0.758562 5.8759036 0.641689 4.4710843 15.44145 0.7493976 4.431629 14.766265 1.885737 4.7891566 6.89088 7.560241 45.03798 34.722892 0.275812 0.8433133 6.824598 6.560241 67.93058 52.662651 6.745719 27.274096 4.889888 14.471386 94.41713 72.174699 81.02243 67.572289 5.795205 20.003675 0.079416 0.5005422 0.016391 0.0453615 21.25847 17.277108 53.24314 97.849398 9.739514 10.120482 0.026056 1.0078313 34.648434 6.534348 -7.6283133 1.321913 61.46583 0.392521 0.147262 0.254344 0.233483 1.19335 69.54238 0.9884337 0.4084337 0.6390964 1.5896988 -0.0086145 142.39157 112 N 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 166 Hind Limb F R209H Mean 0.08846591 35.93068182 0.03895455 15.58409091 0.12122727 48.48409091 0.16014773 64.06931818 0.24863636 24.28409091 75.71590909 1.82727273 5.90568182 4.21590909 0.45227273 15.65681818 5.24431818 8.19318182 37.97727273 0.97397727 6.97727273 53.84090909 25.94886364 16.61125 79.59090909 72 22.93125 0.49590909 0.05272727 19.07954545 101.2954546 10.76136364 1.01545455 34.42875 -7.84977273 109.4545455 1.23068182 0.44352273 0.72954545 1.69352273 -0.02136364 99.01136364 SD 0.010834 4.442386 0.012517 4.255784 0.025295 5.061345 0.030727 4.442386 0.035716 6.276995 6.276995 0.36632 0.864811 0.598059 16.46886 4.845256 2.029884 7.600451 53.14521 0.48285 7.549039 68.31551 5.887372 7.523195 97.67394 86.10726 9.096783 0.084386 0.024947 23.70665 56.13001 10.67169 0.055974 6.45687 1.00655 64.65877 0.543744 0.197449 0.331666 0.296365 1.27859 70.34978 N 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 88 FDR Yes No No No Yes Yes No No No Yes Yes No No Yes No No No No No No No No No No No No Yes No No No No No No No No No Yes No No Yes No Yes 178! Table S3.7 Gait analysis parameters of male Gnao1 KO mutant mice Fore Limb Hind Limb FDR Yes No No Yes No No No No No No No No Yes No No No No No No Yes No No No No No No No P value <0.000001 0.000566 0.72381 0.098909 0.549673 0.030644 0.662124 0.000566 0.036704 0.430085 0.430085 0.000456 0.000016 0.029354 0.929857 0.000132 0.975957 0.968237 0.887753 0.001399 0.86119 0.083955 0.970756 0.1875 0.611842 0.668913 0.001224 FDR Yes Yes No No No No No Yes No No No Yes Yes No No Yes No No No Yes No No No No No No Yes 0.055625 0.032442 0.974929 0.558834 0.277738 0.220542 0.282875 0.049741 0.236915 0.001913 0.840325 No No No No No No No No No No No 0.067872 0.000027 0.526683 0.010604 0.843972 0.558834 0.036413 0.000003 0.111301 0.282875 0.049741 0.394142 <0.000001 0.817383 0.000026 No Yes No No No No No Yes No No No No Yes No Yes P value 0.000054 0.024865 0.165587 0.000746 0.011992 0.075845 0.224409 0.024765 0.017583 0.003525 0.003525 0.007551 0.000016 0.010131 0.4019 0.117131 0.033279 0.098559 0.582856 0.000188 0.954784 0.90312 0.908455 0.229963 0.57553 0.969952 0.041877 Measured Parameters X.SwingStride X.BrakeStride X.PropelStride X.StanceStride Swing Brake Propel Stance Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle StanceWidth StepAngle SLVar SWVar StepAngleVar X.Steps Stride.Length.CV Stance.Width.CV Step.Angle.CV Absolute.PawAngle Paw.Angle.Variability Swing.Duration.CV Paw.Area.at.Peak.Stanc e.in.sq..cm Paw.Area.Variability.at.P eak.Stan Hind.Limb.Shared.Stan ce.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositioni ng.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag M WT Mean 0.0966143 38.779286 0.0560429 22.397857 0.0979429 38.821429 0.1540143 61.220714 0.2506429 36.537143 63.462857 1.5964286 6.29 4.185 -0.705714 4.9457143 7.0635714 3.45 61.464286 1.4037143 13.928571 58.392857 21.817857 22.860071 46.407143 54.4 28.250429 SD 0.013541 3.410721 0.014514 4.576582 0.022044 4.457018 0.028645 3.410721 0.03881 6.893364 6.893364 0.227723 0.818869 0.630966 5.816768 3.114537 2.029963 2.661543 66.30954 0.373751 14.98519 74.79071 4.385313 7.555922 53.94267 73.56002 9.041509 0.2716429 0.051797 0.021 0.008591 0 0 0 1 1 1 8.7571429 1.0077143 16.484857 -4.639 9.097293 0.051682 2.680563 0.926215 1.3222857 0.336865 0.4799286 0.9550714 -2.694786 0.0115 1 0.173721 0.340194 0.350606 0.801008 0 SD 0.013372 4.330104 0.015854 5.745158 0.021646 5.447477 0.027824 4.331596 0.035769 8.483339 8.483339 0.304477 0.777601 0.644087 5.154065 3.095935 2.431709 2.092217 61.51358 0.305636 15.00567 73.8727 4.550977 6.790976 56.61555 72.00602 7.979169 0.055659 0.006557 0 0 0 8.903614 0.025839 3.247951 1.002793 0.340982 0.215823 0.333297 0.389434 0.768448 0 N 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 M WT Mean 0.09272143 37.1207143 0.03049286 12.1007143 0.12967857 50.775 0.16015 62.8792857 0.25293571 19.2214286 80.7785714 1.735 6.33285714 4.16285714 0.34071429 18.8221429 4.76857143 7.68571429 30.8785714 1.13464286 8.29285714 59.25 21.6285714 18.1392143 72.3 59.2642857 24.3625714 SD 0.010933 4.462917 0.00852 2.923768 0.029871 4.326131 0.034105 4.462917 0.040235 4.247196 4.247196 0.333269 0.759443 0.663579 19.33209 4.126107 1.893194 6.825772 42.43293 0.419421 10.34269 74.44153 4.351206 7.026218 91.3001 72.92655 8.268823 0.59078571 0.059483 0.05207143 0.027261 17.5357143 87.6285714 10.3857143 1.00771429 43.0139286 -8.8659286 110 1.32228571 0.47992857 0.77992857 1.45178571 -0.0047143 103.528571 21.94513 49.65526 10.55481 0.051682 4.791189 1.463116 64.09716 0.336865 0.173721 0.325529 0.368579 1.32741 66.43366 Hind Limb M KO Mean 0.08395 35.062 0.03093 12.858 0.12735 52.082 0.15824 64.938 0.24218 19.738 80.262 1.899 5.899 4.352 0.127 16.617 4.761 7.65 31.66 0.9808 8.54 43.77 21.65 17.0179 66.41 63.41 21.0523 SD 0.010752 4.550764 0.010592 4.159477 0.029423 4.937619 0.032238 4.550764 0.037453 5.879957 5.879957 0.377524 0.740665 0.652653 17.32487 4.610157 1.949302 6.867248 41.96396 0.265626 11.37285 58.11359 4.60758 5.621839 84.49746 75.36405 6.891723 0.607 0.077401 0.0391 0.015641 19.41 104.62 10.12 1.011 44.5698 -7.9607 123.91 1.37 0.5296 0.7451 1.0712 0.035 139.88 23.43919 51.38344 9.93136 0.025839 6.666559 1.404932 69.66487 0.340982 0.215823 0.290925 0.43125 1.290374 62.16779 N 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 N 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 Fore Limb M KO Mean 0.08936 37.65 0.05879 24.677 0.09069 37.679 0.1495 62.351 0.23888 39.468 60.532 1.689 5.828 4.401 -1.316 4.306 7.682 2.92 56.83 1.2313 14.04 59.58 21.885 21.718 50.45 54.04 25.9427 0.2851 0.0188 1 1 8.72 1.011 16.9022 -4.4848 1 1.37 0.5296 0.9027 -2.8457 0.0323 1 N 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 140 179! Table S3.8 Gait analysis parameters of female Gnao1 KO mutant mice N 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 M WT Mean 0.08811111 36.402381 0.03019841 12.4555556 0.12659524 51.1444444 0.15688095 63.5976191 0.24495238 19.5436508 80.4571429 1.78571429 6.0047619 4.31825397 -0.3563492 16.1230159 4.68571429 7.17460317 31.4126984 1.10555556 8.01587302 47.5238095 23.5277778 18.5309524 52.7698413 47.9444444 24.2536508 SD 0.011389 4.265311 0.009046 3.618786 0.030455 4.588133 0.033709 4.265311 0.040905 5.338532 5.338351 0.327344 0.663669 0.718265 16.56968 3.556811 1.67345 6.33603 41.97052 0.359134 9.890589 61.83885 4.458836 5.962569 74.45967 60.44056 6.894116 0.58753968 0.118644 0.05039683 0.021887 18.1746032 82.8174603 9.76984127 1.01571429 42.4457143 -8.7710318 93.0396825 1.31436508 0.39960317 0.82809524 1.37214286 -0.026746 96.7857143 21.7576 44.47514 9.806253 0.04399 8.392421 2.194612 54.28361 0.442389 0.21485 0.318601 0.462222 1.208638 59.58625 N 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 Hind Limb M KO Mean 0.08773684 36.98947368 0.03043421 12.71184211 0.12134211 50.30657895 0.15181579 63.01052632 0.23951316 20.11447368 79.88552632 1.73552632 5.77763158 4.41315789 1.41184211 15.79342105 6.30394737 6.96052632 33.48684211 1.13855263 11.75 46.46052632 24.89473684 19.94105263 52.28947368 52.36842105 25.08039474 SD 0.011287 3.859001 0.0087 3.000665 0.026067 3.779959 0.030911 3.859001 0.038918 4.377875 4.377875 0.279859 0.581171 0.723573 16.49608 4.62412 4.855964 6.163204 44.48086 0.330473 15.73203 59.4231 5.08548 6.327376 70.78961 63.93452 7.679944 0.57828947 0.068729 0.05644737 0.031271 16.5 73.63157895 9.60526316 1.00421053 42.05407895 -8.63855263 100.7105263 1.19828947 0.41618421 0.90736842 1.26828947 -0.00013158 110.8289474 19.89874 38.92106 9.929188 0.050153 4.853255 1.626896 53.62383 0.412137 0.290181 0.298098 0.288543 1.174256 54.61755 N 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 76 1.3143651 0.442389 1.19828947 0.412137 Fore Limb Hind Limb P value 0.80571 0.727928 0.959871 0.766918 0.433018 0.611446 0.595756 0.727928 0.63591 0.68239 0.68239 0.692429 0.085256 0.630716 0.013715 0.044961 0.007312 0.297271 0.489835 0.565803 0.444068 0.705629 0.139258 0.955387 0.330411 0.612745 0.338005 FDR No No No No No No No No No No No No No No No No No No No No No No No No No No No P value 0.820639 0.327439 0.855721 0.604358 0.212013 0.18154 0.287305 0.327439 0.352323 0.432768 0.432118 0.26692 0.014479 0.365384 0.462617 0.570224 0.00075 0.814438 0.739738 0.515402 0.039537 0.904505 0.046748 0.113165 0.963959 0.622492 0.430043 FDR No No No No No No No No No No No No No No No No No No No No No No No No No No No 0.040811 0.055277 0.175091 0.089351 0.088329 0.25101 0.06535 0.642852 0.406479 0.306676 0.630299 No No No No No No No No No No No 0.536284 0.108075 0.58501 0.138078 0.908552 0.089351 0.711101 0.648945 0.329558 0.06535 0.642852 0.080856 0.079655 0.878369 0.095768 No No No No No No No No No No No No No No No Measured Parameters X.SwingStride X.BrakeStride X.PropelStride X.StanceStride Swing Brake Propel Stance Stride X.BrakeStance X.PropelStance Stance.Swing StrideLength Stride.Frequency PawAngle StanceWidth StepAngle SLVar SWVar StepAngleVar X.Steps Stride.Length.CV Stance.Width.CV Step.Angle.CV Absolute.PawAngle Paw.Angle.Variability Swing.Duration.CV Paw.Area.at.Peak.Stanc e.in.sq..cm Paw.Area.Variability.at.P eak.Stan Hind.Limb.Shared.Stan ce.Time X..Shared.Stance StanceFactor Gait.Symmetry MAX.dA.dT MIN.dA.dT Tau..Propulsion Overlap.Distance PawPlacementPositioni ng.PPP. Ataxia.Coefficient Midline.Distance Axis.Distance Paw.Drag M WT Mean 0.090881 37.835714 0.0605079 25.046825 0.0901111 37.116667 0.1505952 62.164286 0.2415397 40.286508 59.713492 1.6603175 5.9293651 4.3777778 1.4746032 4.2285714 6.8833333 3.7857143 54.34127 1.3600794 14.293651 49.730159 23.960317 23.392857 46.468254 47.706349 28.772381 SD 0.015084 3.303319 0.016707 4.950098 0.022099 5.123838 0.02981 3.303319 0.041883 7.425463 7.425463 0.239359 0.731253 0.727174 5.189311 3.331207 1.681987 2.986924 59.0929 0.343405 15.47285 62.78731 4.61762 6.995052 51.84063 61.74529 7.820674 0.2765079 0.054792 0.0228571 0.010648 1 1 1 10.134921 1.0157143 16.179365 -4.832778 10.84019 0.04399 3.068414 1.217865 0 0 0 0.3996032 1.0359524 -2.553175 0.013254 1 0.21485 0.359989 0.490671 0.758111 0 N 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 126 Fore Limb M KO Mean 0.09034211 37.99342105 0.06063158 25.26184211 0.08768421 36.73684211 0.14834211 62.00657895 0.23867105 40.73947368 59.26052632 1.64736842 5.75657895 4.42894737 -0.54342105 5.20657895 7.89210526 3.35526316 48.53947368 1.3325 16.11842105 53.19736842 24.99342105 23.44684211 54.17105263 52.38157895 27.70763158 SD 0.015029 2.778601 0.017212 5.049963 0.01981 5.16622 0.02814 2.778601 0.041275 7.909034 7.909034 0.19898 0.608678 0.739336 6.194951 3.347649 3.577812 2.564912 55.40865 0.306725 17.7981 63.64171 5.068526 5.989497 58.31349 66.31239 7.309925 0.26118421 0.04472 0.02763158 0.024214 1 1 1 8.15789474 1.00421053 15.49092105 -4.62289474 8.423859 0.050153 2.176046 1.315192 0 0 0 0.41618421 1.07894737 -2.48618421 0.06552632 1 0.290181 0.348879 0.372673 0.727057 0 180! Feng%at%al%Table&S5& &Benchling)off,target)list)for)Gnao1)G203)gRNA)) & Table S3.9 Benchling off-target list for Gnao1 G203R gRNA PAM& Score& Chromosome& Mismatches& 0% 4) 4) 2) 4) 4) 4) 4) 4) 3) 4) 4) 4) 4) 4) 4) 4) 3) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 3) 4) GGG% Sequence& TGCAGGCTGTTTGACGTCGG% GGCAAGCTGATTGACGTCTG) TAG) TGGATGGTGTTGGACGTCGG) AAG) TGCAGGCTGTTTGAAGTCTG) CAG) GGTGGGCTGTTTGACGTGGG) AGG) TTCAGGCTGAGTGACGTCAG) TGG) AGCAGGCACTTTGAAGTCGG) AAG) TTCAGTCTGTTAGACGTCTG) TAG) TGCATGGGGTTTGACTTCGG) AGG) TGCTGGCTGTTTGAGGTGGG) AAG) TCCAGGCTGGTGGACGTGGG) CAG) TGATGGCTGTTCGACTTCGG) GAG) TACAGAATGTTTGACGTGGG) AGG) TTCAGTCTGTTTGAGGTCGT) TGG) AGCAGGCTGCTTGACATCGA) GAG) TGCAAGCTGGTTGAGGTCAG) GGG) TCCAGGATGTTTGATGCCGG) AAG) TGCAGGCTGTCTGAAGTCTG) GGG) GGCTGGCTGTTTGACCTCAG) AGG) AGCAGCCTGTTTGAAGTCTG) TGG) GGCAGGCTGTATGAAGGCGG) AGG) TGGAGGCTGTTACACGTCAG) CAG) TGCTGGCTATTTGAAGTCTG) AGG) TGCTGGTTATTTGTCGTCGG) GAG) TCCAGGCTGTCTGATGTCAG) GAG) TTCAGGATGTTTGACGTATG) CAG) TGCACGCTGTGAGACGTGGG) CGG) TGCATGCTGTCTGAAGTCAG) AAG) TGCAGGCTGTATGACCTCTG) GGG) TGCAGTCTCTTTGACGACAG) TGG) Gene& chr8% 100% ENSMUSG00000031748% ) chr18) 0.6189) 0.5200) ENSMUSG00000041390) chr6) ) chr3) 0.5076) 0.3804) ) chr1) 0.3169) ENSMUSG00000032497) chr9) ) chr3) 0.2931) 0.1953) ) chr1) 0.1929) ) chr13) ) chr1) 0.1923) 0.1710) ) chr1) 0.1556) ENSMUSG00000086805) chr8) 0.1543) ENSMUSG00000057614) chr5) 0.1515) ) chrX) ) chr4) 0.1480) 0.1450) ) chr17) ) chr18) 0.1403) 0.1343) ENSMUSG00000026413) chr1) 0.1262) ) chrX) ) chr11) 0.1144) 0.1127) ) chr5) ) chr1) 0.1127) 0.1004) ) chr10) 0.1002) ) chr11) ) chrX) 0.0954) 0.0933) ) chr3) 0.0930) ENSMUSG00000020015) chr10) 0.0865) ) chrX) 0.0862) ) chr2) ) chr11) 0.0836) 181! Table S3.9 (cont’d) TGCATGCTGTAGGACCTCGG) AGG) CGGAGGCTGTTTGACTTGGG) AGG) TCCAGGCTGTTTCAGGACGG) AAG) GGCAGCCTGTTTGACATCAG) GAG) TGCAAGATGTTTGACCTCAG) AAG) TGGAGGTTGTTTGAGGTAGG) AGG) TGTGGGCTGTTTGACCTGGG) AGG) GGCAGGCTGTTTGAAGCCAG) GGG) TCCAGGCTGTTTGAGGGCTG) CAG) TGCAGGCTGGCTGACGATGG) TGG) TGCAGGATGCTTGACCTCTG) TAG) TGCACTCTGTTTGAGGTTGG) AGG) TCCAGGCTGTGTGAGGTGGG) AGG) GGCAGGCTGTTGGAAGTAGG) GAG) TGAAGGCTGTTCGAAGTGGG) GAG) TGCAGGCTGATTGATGGCTG) GAG) TACAGACTGTTTGACTTGGG) CAG) TTCAGGCTGTTTTACTTCTG) AGG) AGCAGGATGTTTGTCGTGGG) GAG) AGCAGGCTGTGTGACCTGGG) AGG) ) chr4) 0.0771) ) chr5) 0.0754) ) chr8) 0.0745) ) chr17) 0.0744) ) chr19) 0.0720) ) chr2) 0.0704) ) chr19) 0.0693) ) chr9) 0.0669) 0.0647) ENSMUSG00000097637) chr8) ) chr8) 0.0611) 0.0604) ) chr2) 0.0599) ) chr10) ) chr9) 0.0574) 0.0520) ) chr2) ) chrX) 0.0480) 0.0470) ) chr7) 0.0430) ) chr3) ) chr15) 0.0422) 0.0420) ) chr1) ) chr6) 0.0385) )))))))))))))))))))))Row)1)includes)the)on,target)gRNA)for)the)Gnao1)G203)site.)Off,target)hits)are)scored)and)ranked)by)an) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) 4) inverse)likelihood)of)off,target)binding.)If)an)off,target)is)predicted)to)occur)within)a)coding)region)of)a) gene,)the)Ensembl)number)of)the)affected)locus)is)listed)in)the)Gene)column.)Analysis)was)performed)on) the)Benchling)platform)using)reference)genome)GRCM38)(MM10,)Mus)Musculus),)guide)length)of)20bp,) and)an)NGG)PAM.) Row 1 includes the on-target gRNA for the Gnao1 G203 site. Off-target hits are scored and ranked by an inverse likelihood of off-target binding. If an off-target is predicted to occur within a coding region of a gene, the Ensembl number of the affected locus is listed in the Gene column. Analysis was performed on the Benchling platform using reference genome GRCM38 (MM10, Mus Musculus), guide length of 20bp, and an NGG PAM. 182! REFERENCES 183! REFERENCES Ananth, A. L., Robichaux-Viehoever, A., Kim, Y. M., Hanson-Kahn, A., Cox, R., Enns, G. M., . . . Bernstein, J. A. (2016). Clinical Course of Six Children With GNAO1 Mutations Causing a Severe and Distinctive Movement Disorder. Pediatr Neurol, 59, 81-84. doi:10.1016/j.pediatrneurol.2016.02.018 Arya, R., Spaeth, C., Gilbert, D. L., Leach, J. L., & Holland, K. D. (2017). GNAO1-associated epileptic encephalopathy and movement disorders: c.607G>A variant represents a probable mutation hotspot with a distinct phenotype. Epileptic Disord, 19(1), 67-75. doi:10.1684/epd.2017.0888 Blumkin, L., Lerman-Sagie, T., Westenberger, A., Ben-Pazi, H., Zerem, A., Yosovich, K., & Lev, D. (2018). Multiple Causes of Pediatric Early Onset Chorea-Clinical and Genetic Approach. Neuropediatrics, 49(4), 246-255. doi:10.1055/s-0038-1645884 Bromberg, K. D., Iyengar, R., & He, J. C. (2008). Regulation of neurite outgrowth by G(i/o) signaling pathways. Front Biosci, 13, 4544-4557. Bruun, T. U. J., DesRoches, C. L., Wilson, D., Chau, V., Nakagawa, T., Yamasaki, M., . . . Mercimek-Andrews, S. (2018). Prospective cohort study for identification of underlying genetic causes in neonatal encephalopathy using whole-exome sequencing. Genet Med, 20(5), 486-494. doi:10.1038/gim.2017.129 Danti, F. R., Galosi, S., Romani, M., Montomoli, M., Carss, K. J., Raymond, F. L., . . . Guerrini, R. (2017). GNAO1 encephalopathy: Broadening the phenotype and evaluating e143. doi:10.1212/NXG.0000000000000143 outcome. Neurol Genet, treatment 3(2), and Deacon, R. M. (2013). Measuring the strength of mice. J Vis Exp(76). doi:10.3791/2610 Dhamija, R., Mink, J. W., Shah, B. B., & Goodkin, H. P. (2016). GNAO1-Associated 615-617. Disorder. Mov Disord Pract, Movement doi:10.1002/mdc3.12344 Clin 3(6), Dhamija, R. M., J, W.; Shah, B, B.; Goodkin, H, P. (2016). GNAO1-Associated Movement Disorder. Movement Disorders Clinical Practice, 3(6), 615-617. 184! doi:10.1002/mdc3.12344 Dhir, A. (2012). Pentylenetetrazol (PTZ) kindling model of epilepsy. Curr Protoc Neurosci, Chapter 9, Unit9.37. doi:10.1002/0471142301.ns0937s58 DiBello, P. R., Garrison, T. R., Apanovitch, D. M., Hoffman, G., Shuey, D. J., Mason, K., . . . Dohlman, H. G. (1998). Selective uncoupling of RGS action by a single point mutation in the G protein alpha-subunit. J Biol Chem, 273(10), 5780-5784. Dietel, T. H., T. B.; Schlüter, G.; Cordes, I.; Trollmann, R.; Bast, T. (2016). Genotype and phenotype in GNAO1-Mutation – Case report of an unusual course of a childhood epilepsy. Neuropediatrics, 47(S 01), 1-17. doi:10.1055/s-0036-1583625 Doench, J. G., Fusi, N., Sullender, M., Hegde, M., Vaimberg, E. W., Donovan, K. F., . . . Root, D. E. (2016). Optimized sgRNA design to maximize activity and minimize off-target effects of CRISPR-Cas9. Nat Biotechnol, 34(2), 184-191. doi:10.1038/nbt.3437 Euro, E.-R. E. S. C., Epilepsy Phenome/Genome, P., & Epi, K. C. (2014). De novo mutations in synaptic transmission genes including DNM1 cause epileptic encephalopathies. 360-370. doi:10.1016/j.ajhg.2014.08.013 Genet, 95(4), Hum Am J Feng, H., Khalil, S., Neubig, R. R., & Sidiropoulos, C. (2018). A mechanistic review on Dis. movement Neurobiol disorder. GNAO1-associated doi:10.1016/j.nbd.2018.05.005 Feng, H., Larrivee, C. L., Demireva, E. Y., Xie, H., Leipprandt, J. R., & Neubig, R. R. (2019). Mouse models of GNAO1-associated movement disorder: Allele- and sex-specific differences in phenotypes. PLoS One, 14(1), e0211066. doi:10.1371/journal.pone.0211066 Feng, H., Sjogren, B., Karaj, B., Shaw, V., Gezer, A., & Neubig, R. R. (2017). Movement disorder in GNAO1 encephalopathy associated with gain-of-function mutations. Neurology, 89(8), 762-770. doi:10.1212/WNL.0000000000004262 Franco-Pons, N., Torrente, M., Colomina, M. T., & Vilella, E. (2007). Behavioral deficits in the cuprizone-induced murine model of demyelination/remyelination. Toxicol Lett, 169(3), 205-213. doi:10.1016/j.toxlet.2007.01.010 185! Franek, M., Pagano, A., Kaupmann, K., Bettler, B., Pin, J. P., & Blahos, J. (1999). The heteromeric GABA-B receptor recognizes G-protein alpha subunit C-termini. Neuropharmacology, 38(11), 1657-1666. Fu, Y., Zhong, H., Nanamori, M., Mortensen, R. M., Huang, X., Lan, K., & Neubig, R. R. (2004). RGS-insensitive G-protein mutations to study the role of endogenous RGS 229-243. Methods doi:10.1016/S0076-6879(04)89014-1 Enzymol, proteins. 389, Gawlinski, P., Posmyk, R., Gambin, T., Sielicka, D., Chorazy, M., Nowakowska, B., . . . Wiszniewski, W. (2016). PEHO Syndrome May Represent Phenotypic Expansion at the Severe End of the Early-Onset Encephalopathies. Pediatr Neurol, 60, 83-87. doi:10.1016/j.pediatrneurol.2016.03.011 Gazi, L., Nickolls, S. A., & Strange, P. G. (2003). Functional coupling of the human dopamine D2 receptor with G alpha i1, G alpha i2, G alpha i3 and G alpha o G proteins: evidence for agonist regulation of G protein selectivity. Br J Pharmacol, 138(5), 775-786. doi:10.1038/sj.bjp.0705116 Gillies, G. E., Murray, H. E., Dexter, D., & McArthur, S. (2004). Sex dimorphisms in the neuroprotective effects of estrogen in an animal model of Parkinson's disease. Pharmacol Biochem Behav, 78(3), 513-522. doi:10.1016/j.pbb.2004.04.022 Goldenstein, B. L., Nelson, B. W., Xu, K., Luger, E. J., Pribula, J. A., Wald, J. M., . . . Doze, V. A. (2009). Regulator of G protein signaling protein suppression of Galphao protein-mediated alpha2A adrenergic receptor inhibition of mouse hippocampal CA3 epileptiform activity. Mol Pharmacol, 75(5), 1222-1230. doi:mol.108.054296 [pii]10.1124/mol.108.054296 Grecksch, G., Becker, A., Schroeder, H., Kraus, J., Loh, H., & Hollt, V. (2004). Accelerated kindling development in mu-opioid receptor deficient mice. Naunyn 287-293. Schmiedebergs doi:10.1007/s00210-004-0870-4 Pharmacol, 369(3), Arch Haeussler M, S. K., Eckert H, Eschstruth A, Mianné J, Renaud JB, Schneider-Maunoury S, Shkumatava A, Teboul L, Kent J, Joly JS, Concordet JP. (2016). Evaluation of off-target and on-target scoring algorithms and integration into the guide RNA selection 148. doi:https://doi.org/10.1186/s13059-016-1012-2 CRISPOR. Genome 17(1), tool Biol, 186! Hansen, S. T., & Pulst, S. M. (2013). Response to ethanol induced ataxia between C57BL/6J and 129X1/SvJ mouse strains using a treadmill based assay. Pharmacol Biochem Behav, 103(3), 582-588. doi:10.1016/j.pbb.2012.10.010 Hirata, H., Takahashi, A., Shimoda, Y., & Koide, T. (2016). Caspr3-Deficient Mice Exhibit Low Motor Learning during the Early Phase of the Accelerated Rotarod Task. PLoS One, 11(1), e0147887. doi:10.1371/journal.pone.0147887 Honey, C. M., Malhotra, A. K., Tarailo-Graovac, M., van Karnebeek, C. D. M., Horvath, G., & Sulistyanto, A. (2018). GNAO1 Mutation-Induced Pediatric Dystonic Storm Rescue With Pallidal Deep Brain Stimulation. J Child Neurol, 33(6), 413-416. doi:10.1177/0883073818756134 Hsu, P. D., Scott, D. A., Weinstein, J. A., Ran, F. A., Konermann, S., Agarwala, V., . . . Zhang, F. (2013). DNA targeting specificity of RNA-guided Cas9 nucleases. Nat Biotechnol, 31(9), 827-832. doi:10.1038/nbt.2647 Iyer, V., Boroviak, K., Thomas, M., Doe, B., Riva, L., Ryder, E., & Adams, D. J. (2018). No unexpected CRISPR-Cas9 off-target activity revealed by trio sequencing of gene-edited e1007503. doi:10.1371/journal.pgen.1007503 Genet, 14(7), PLoS mice. Jiang, M., & Bajpayee, N. S. (2009). Molecular mechanisms of go signaling. Neurosignals, 17(1), 23-41. doi:10.1159/000186688 Jiang, M., Gold, M. S., Boulay, G., Spicher, K., Peyton, M., Brabet, P., . . . Birnbaumer, L. (1998). Multiple neurological abnormalities in mice deficient in the G protein Go. Proc Natl Acad Sci U S A, 95(6), 3269-3274. doi:10.1073/pnas.95.6.3269 Kehrl, J. M., Sahaya, K., Dalton, H. M., Charbeneau, R. A., Kohut, K. T., Gilbert, K., . . . Neubig, R. R. (2014). Gain-of-function mutation in Gnao1: a murine model of epileptiform encephalopathy (EIEE17)? Mamm Genome, 25(5-6), 202-210. doi:10.1007/s00335-014-9509-z Kelly, M., Park, M., Mihalek, I., Rochtus, A., Gramm, M., Perez-Palma, E., . . . Poduri, A. (2019). Spectrum of neurodevelopmental disease associated with the GNAO1 guanosine 406-418. doi:10.1111/epi.14653 triphosphate-binding Epilepsia, region. 60(3), Kompoliti, K. (1999). Estrogen and movement disorders. Clin Neuropharmacol, 22(6), 187! 318-326. Kulkarni, N., Tang, S., Bhardwaj, R., Bernes, S., & Grebe, T. A. (2016). Progressive Movement Disorder in Brothers Carrying a GNAO1 Mutation Responsive to Deep Brain 211-214. doi:10.1177/0883073815587945 Stimulation. 31(2), J Child Neurol, Lan, K. L., Sarvazyan, N. A., Taussig, R., Mackenzie, R. G., DiBello, P. R., Dohlman, H. G., & Neubig, R. R. (1998). A point mutation in Galphao and Galphai1 blocks interaction with regulator of G protein signaling proteins. J Biol Chem, 273(21), 12794-12797. Larrivee, C. L., Feng, H., Leipprandt, J. R., Demireva, E. Y., Xie, H., & Neubig, R. R. (2019). Mice with GNAO1 R209H Movement Disorder Variant Display Hyperlocomotion Alleviated by Risperidone. bioRxiv. Law, C. Y., Chang, S. T., Cho, S. Y., Yau, E. K., Ng, G. S., Fong, N. C., & Lam, C. W. (2015). Clinical whole-exome sequencing reveals a novel missense pathogenic variant of GNAO1 in a patient with infantile-onset epilepsy. Clin Chim Acta, 451(Pt B), 292-296. doi:10.1016/j.cca.2015.10.011 Li, Q., Lau, A., Morris, T. J., Guo, L., Fordyce, C. B., & Stanley, E. F. (2004). A syntaxin 1, Galpha(o), and N-type calcium channel complex at a presynaptic nerve terminal: analysis by quantitative immunocolocalization. J Neurosci, 24(16), 4070-4081. doi:10.1523/JNEUROSCI.0346-04.2004 Lorenzen, A., Lang, H., & Schwabe, U. (1998). Activation of various subtypes of G-protein alpha subunits by partial agonists of the adenosine A1 receptor. Biochem Pharmacol, 56(10), 1287-1293. Madalan, A., Yang, X., Ferris, J., Zhang, S., & Roman, G. (2012). G(o) activation is required for both appetitive and aversive memory acquisition in Drosophila. Learn Mem, 19(1), 26-34. doi:10.1101/lm.024802.111 Marce-Grau, A., Dalton, J., Lopez-Pison, J., Garcia-Jimenez, M. C., Monge-Galindo, L., Cuenca-Leon, E., . . . Macaya, A. (2016). GNAO1 encephalopathy: further delineation of a severe neurodevelopmental syndrome affecting females. Orphanet J Rare Dis, 11, 38. doi:10.1186/s13023-016-0416-0 Marecos, C., Duarte, S., Alonso, I., Calado, E., & Moreira, A. (2018). [GNAO1: a new 188! gene to consider on early-onset childhood dystonia]. Rev Neurol, 66(9), 321-322. Menke, L. A., Engelen, M., Alders, M., Odekerken, V. J., Baas, F., & Cobben, J. M. (2016). Recurrent GNAO1 Mutations Associated With Developmental Delay and a Movement 1598-1601. doi:10.1177/0883073816666474 Disorder. Neurol, 31(14), J Child Nakamura, K., Kodera, H., Akita, T., Shiina, M., Kato, M., Hoshino, H., . . . Saitsu, H. (2013). De Novo mutations in GNAO1, encoding a Galphao subunit of heterotrimeric G proteins, cause epileptic encephalopathy. Am J Hum Genet, 93(3), 496-505. doi:10.1016/j.ajhg.2013.07.014 Oleas, J., Yokoi, F., DeAndrade, M. P., Pisani, A., & Li, Y. (2013). Engineering animal models of dystonia. Mov Disord, 28(7), 990-1000. doi:10.1002/mds.25583 Parr, T., & Friston, K. J. (2017). Working memory, attention, and salience in active inference. Sci Rep, 7(1), 14678. doi:10.1038/s41598-017-15249-0 Qin, W., Dion, S. L., Kutny, P. M., Zhang, Y., Cheng, A. W., Jillette, N. L., . . . Wang, H. (2015). Efficient CRISPR/Cas9-Mediated Genome Editing in Mice by Zygote Electroporation 423-430. doi:10.1534/genetics.115.176594 Nuclease. Genetics, 200(2), of Revankar, C. M., Cimino, D. F., Sklar, L. A., Arterburn, J. B., & Prossnitz, E. R. (2005). A transmembrane intracellular estrogen receptor mediates rapid cell signaling. Science, 307(5715), 1625-1630. doi:10.1126/science.1106943 Saitsu, H., Fukai, R., Ben-Zeev, B., Sakai, Y., Mimaki, M., Okamoto, N., . . . Matsumoto, N. (2016). Phenotypic spectrum of GNAO1 variants: epileptic encephalopathy to involuntary movements with severe developmental delay. Eur J Hum Genet, 24(1), 129-134. doi:10.1038/ejhg.2015.92 Schirinzi, T., Garone, G., Travaglini, L., Vasco, G., Galosi, S., Rios, L., . . . Leuzzi, V. (2019). Phenomenology and clinical course of movement disorder in GNAO1 variants: Results from an analytical review. Parkinsonism Relat Disord, 61, 19-25. doi:10.1016/j.parkreldis.2018.11.019 Schorling, D. C., Dietel, T., Evers, C., Hinderhofer, K., Korinthenberg, R., Ezzo, D., . . . Kirschner, J. (2017). Expanding Phenotype of De Novo Mutations in GNAO1: Four New Cases and Review of Literature. Neuropediatrics, 48(5), 371-377. 189! doi:10.1055/s-0037-1603977 Schutsky, K., Ouyang, M., & Thomas, S. A. impairs hippocampus-dependent emotional memory via Gi/o-coupled beta2-adrenergic signaling. Learn Mem, 18(9), 598-604. doi:10.1101/lm.2302811 (2011). Xamoterol retrieval Smith, K. M., & Dahodwala, N. (2014). Sex differences in Parkinson's disease and other 44-56. Neurol, 259, movement doi:10.1016/j.expneurol.2014.03.010 disorders. Exp Solis, G. P., & Katanaev, V. L. (2018). Galphao (GNAO1) encephalopathies: plasma 23846-23847. functions. Oncotarget, 9(35), membrane doi:10.18632/oncotarget.22067 vs. Golgi Song, C. H., Fan, X., Exeter, C. J., Hess, E. J., & Jinnah, H. A. (2012). Functional analysis of dopaminergic systems in a DYT1 knock-in mouse model of dystonia. Neurobiol Dis, 48(1), 66-78. doi:10.1016/j.nbd.2012.05.009 Strittmatter, S. M., Fishman, M. C., & Zhu, X. P. (1994). Activated mutants of the alpha subunit of G(o) promote an increased number of neurites per cell. J Neurosci, 14(4), 2327-2338. Stroobants, S., Gantois, I., Pooters, T., & D'Hooge, R. (2013). Increased gait variability in mice with small cerebellar cortex lesions and normal rotarod performance. Behav Brain Res, 241, 32-37. doi:10.1016/j.bbr.2012.11.034 Tatem, K. S., Quinn, J. L., Phadke, A., Yu, Q., Gordish-Dressman, H., & Nagaraju, K. (2014). Behavioral and locomotor measurements using an open field activity monitoring system for skeletal muscle diseases. J Vis Exp(91), 51785. doi:10.3791/51785 Tian, W. N., Duzic, E., Lanier, S. M., & Deth, R. C. (1994). Determinants of alpha 2-adrenergic receptor activation of G proteins: evidence for a precoupled receptor/G protein state. Mol Pharmacol, 45(3), 524-531. Truett, G. E., Heeger, P., Mynatt, R. L., Truett, A. A., Walker, J. A., & Warman, M. L. (2000). Preparation of PCR-quality mouse genomic DNA with hot sodium hydroxide and tris (HotSHOT). Biotechniques, 29(1), 52, 54. 190! Valenzuela, D., Han, X., Mende, U., Fankhauser, C., Mashimo, H., Huang, P., . . . Fishman, M. C. (1997). G alpha(o) is necessary for muscarinic regulation of Ca2+ channels in mouse heart. Proc Natl Acad Sci U S A, 94(5), 1727-1732. doi:10.1073/pnas.94.5.1727 Waak, M., Mohammad, S. S., Coman, D., Sinclair, K., Copeland, L., Silburn, P., . . . Malone, S. (2018). GNAO1-related movement disorder with life-threatening exacerbations: movement phenomenology and response to DBS. J Neurol Neurosurg Psychiatry, 89(2), 221-222. doi:10.1136/jnnp-2017-315653 Wettschureck, N., & Offermanns, S. (2005). Mammalian G proteins and their cell type 1159-1204. Physiol specific doi:10.1152/physrev.00003.2005 functions. Rev, 85(4), Wilczynski, G. M., Konopacki, F. A., Wilczek, E., Lasiecka, Z., Gorlewicz, A., Michaluk, P., . . . Kaczmarek, L. (2008). Important role of matrix metalloproteinase 9 in epileptogenesis. J Cell Biol, 180(5), 1021-1035. doi:10.1083/jcb.200708213 Wilmshurst, P. T., Byrne, J. C., & Webb-Peploe, M. M. (1989). Relation between interatrial shunts and decompression sickness in divers. Lancet, 2(8675), 1302-1306. Wilson, B. K., & Hess, E. J. (2013). Animal models for dystonia. Mov Disord, 28(7), 982-989. doi:10.1002/mds.25526 Wooten, G. F., Currie, L. J., Bovbjerg, V. E., Lee, J. K., & Patrie, J. (2004). Are men at greater risk for Parkinson's disease than women? J Neurol Neurosurg Psychiatry, 75(4), 637-639. Xiong, J., Peng, J., Duan, H. L., Chen, C., Wang, X. L., Chen, S. M., & Yin, F. (2018). [Recurrent convulsion and pulmonary infection complicated by psychomotor retardation in an infant]. Zhongguo Dang Dai Er Ke Za Zhi, 20(2), 154-157. Yilmaz, S., Turhan, T., Ceylaner, S., Gokben, S., Tekgul, H., & Serdaroglu, G. (2016). Excellent response to deep brain stimulation in a young girl with GNAO1-related progressive 1567-1568. doi:10.1007/s00381-016-3139-6 choreoathetosis. Childs Nerv Syst, 32(9), Zhang, Q., Pacheco, M. A., & Doupnik, C. A. (2002). Gating properties of GIRK channels activated by Galpha(o)- and Galpha(i)-coupled muscarinic m2 receptors in 191! Xenopus oocytes: the role of receptor precoupling in RGS modulation. J Physiol, 545(Pt 2), 355-373. Zhu, X., Petrovski, S., Xie, P., Ruzzo, E. K., Lu, Y. F., McSweeney, K. M., . . . Goldstein, D. B. (2015). Whole-exome sequencing in undiagnosed genetic diseases: interpreting 119 trios. Genet Med, 17(10), 774-781. doi:10.1038/gim.2014.191 192! CHAPTER 4: MICE WITH GNAO1-ASSOCIATED MOVEMENT DISORDER EXHIBIT REDUCED INHIBITORY SYNAPTIC INPUT TO CEREBELLAR PURKINJE CELLS Yukun, Y. did the recording for Figure S4.4 A-H. 193! 4.1 Abstract GNAO1 encodes a heterotrimeric G protein subunit, Gαo, which belongs to the Gi/o family. Mutations in GNAO1 are associated with both early infantile epileptic encephalopathy 17 (EIEE17) and neurodevelopmental disorder with involuntary movement (NEDIM). Our previous finding showed that gain-of-function (GOF) or normal-function (NF) GNAO1 mutations, characterized by their inhibition of cAMP production, are associated with movement disorder patients (Chapter 2) (Feng et al., 2017). The majority of these patients present early onset dystonia or chorea/athetosis, hypotonia, and developmental delay. Although NEDIM patients have been treated with numerous available drugs, few proved to be effective. The pathological mechanisms of this disorder also remain unclear. In this chapter, I provide the first data elucidating neural mechanisms of a human GNAO1 mutant by investigating electrophysiological effects in Gnao1+/G203R mice, a Gnao1-associated movement disorder mouse model. These mice carry one of the most prevalent GNAO1 GOF mutations, G203R. Patch clamp studies of cerebellar Purkinje cells showed significantly lower frequencies of both action potential (AP) related (sIPSCs) and non AP-related (mIPSCs) GABAergic responses in G203R mice. Amplitudes were not affected. Gαo inhibitors reversed this reduction in inhibition significantly, and eliminated the difference between WT and G203R mice. Furthermore, Gαo-coupled α2A adrenergic receptors played a critical role in reducing the sIPSCs while mIPSCs events were regulated by GABAB 194! receptors. The results prove G203R to be a bona fide GOF mutation in an in vivo context, supporting our proposed mechanistic genotype-phenotype correlation of GNAO1-associated neurological disorders. Also, the identification of receptors that regulate both mIPSCs and sIPSCs should facilitate the discovery of new drugs or drug repurposing for GNAO1-associaed disorders. 4.2 Introduction GNAO1 encodes the α subunit of a heterotrimeric G protein, Go, which is the most abundant membrane protein in mammalian central nervous system. It participates in multiple neural signaling pathways (Jiang & Bajpayee, 2009). Mutations in GNAO1 were first found in early onset epileptic encephalopathy (EIEE17). However, since 2016, there have been a growing number of reports on GNAO1 mutation-associated movement disorders with/without seizures. This disorder was officially categorized by OMIM (Online Mendelian Inheritance in Man) in 2017 as neurodevelopmental disorder with involuntary movements (NEDIM). GNAO1-associated NEDIM is a rare neurogenetic disorder, characterized by early onset of hypotonia, movement disorder and developmental delay. Although numerous available drug treatments were tested on NEDIM patients, few proved to be effective (Feng, Khalil, Neubig, & Sidiropoulos, 2018). The pathological mechanisms of this disorder also remained unclear. The Gαo protein functions as a messenger for a broad range of signaling pathways. They include inhibition of cAMP (Levitt, Purington, & Traynor, 2011; Sunahara & Taussig, 195! 2002), inhibition of high-voltage-gated calcium channels (N- and P/Q-type calcium channels) (Ikeda, 1996), and activation of G-protein regulated inward rectifying potassium (GIRK) channels (Zhang, Dickson, & Doupnik, 2004). Our lab has previously established a genotype-phenotype correlation of GNAO1-associated neurological disorders based on Gαo’s canonical pathway of inhibiting cAMP production (Feng et al., 2017). We proposed that loss-of-function (LOF) and partial-loss-of-function (PLOF) GNAO1 mutations are found in epilepsy patients, while the gain-of-function (GOF) and normal-function (NF) mutations are generally found in movement disorder patients (Feng et al., 2017). We have also generated a novel animal model with a knock-in GOF mutation G203R (Feng et al., 2019). Similar to human patients with GNAO1 G203R mutation, this animal model exhibited movement abnormalities in a battery of behavioral assessment including RotaRod, grip strength, and DigiGait (Chapter 3) (Feng et al., 2019). Unlike most GOF GNAO1 mutations, patients with the G203R mutant allele also exhibit early-onset epilepsy (Arya, Spaeth, Gilbert, Leach, & Holland, 2017; Feng et al., 2019; Kelly et al., 2019; Nakamura et al., 2013; Saitsu et al., 2016; Schorling et al., 2017; Xiong et al., 2018), validating the relevance of this animal model, which also showed an increased sensitivity to pentylenetetrazole (PTZ) kindling. Although the G203R animal model mimics the symptoms of human G203R patients, the mechanisms by which these mice develop their movement disorder are still unclear. Also, the GOF nature of the G203R mutation in a physiological environment has only 196! been demonstrated for cAMP regulation in HEK293T cells. It is critical to know if Gαo G203R also has GOF behavior in neurons. To answer these questions, I performed patch clamp studies of Purkinje cells in cerebellar slices of WT and Gnao1+/G203R mutant mice. The cerebellum has long been known for its critical regulation of motor coordination. The evidence for a role in dystonia has begun to emerge. Deficiency in cerebellar motor control may manifest as inaccuracies of visual guided movement, speeded complex movement, loss of muscle tone, abnormal timing, and loss of prediction and coordination (Gowen & Miall, 2007). Clinically, these features are characterized as dysmetria (inaccurate movement), dysdiadochokinesis (inability to execute rapidly alternating movements), hypotonia (reduced muscle tone), and dyscoordination or ataxia (inability to perform smoothly coordinated voluntary movement) (Gowen & Miall, 2007). Structural and/or functional abnormalities of the cerebellum are also associated with dystonia (Bologna & Berardelli, 2018) and chorea (Walker, 2016). Several reports elucidated the role of cerebellum in DYT1 hereditary dystonia (Fremont, Tewari, Angueyra, & Khodakhah, 2017; Song, Bernhard, Hess, & Jinnah, 2014; Vanni et al., 2015). Interestingly, dystonia and chorea/athetosis are the most commonly seen involuntary movements in patients with the G203R mutation (Arya et al., 2017; Kelly et al., 2019; Nakamura et al., 2013; Saitsu et al., 2016; Schorling et al., 2017). Structurally, a core cerebellar circuit mediates all of its function (Eccles, 1967; Reeber, Otis, & Sillitoe, 2013). This circuit centers on Purkinje cells, which are the sole output of the cerebellar 197! cortex (Brown et al., 2019). Purkinje cells receive input from several classes of interneurons. Granule cells project parallel fibers that send excitatory signals to Purkinje cells (Barbour, 1993; Eccles, Llinas, & Sasaki, 1966a, 1966b; Konnerth, Llano, & Armstrong, 1990), while basket cells and stellate cells send inhibitory input to Purkinje cell (Cesana et al., 2013; Hull & Regehr, 2012). In this chapter, I report data on patch clamp recordings of the Purkinje cells in cerebellar slices. I found a decreased frequency in both sIPSC and mIPSC events in G203R mutant mice. A likely mechanism underlying this reduction in frequency is enhanced signaling by the mutant Gαo to mediate presynaptic inhibition of GABA release. Different receptors serve as the driving force for this inhibition. GABAB receptor-activated Gαo mediates non-AP related mIPSCs while α2A adrenergic receptors stimulate AP mediated sIPSCs. Although Gαo-mediated inhibition of high-voltage activated (N, P/Q-type) calcium channels is well-studied (Ikeda, 1996), only α2A adrenergic receptor-mediated sIPSCs function through the activation of membrane-located, voltage-gated calcium channels. GABAB receptor mediated inhibition of mIPSCs is likely to involve Gβγ-mediated direct inhibition of synaptic vesicle release (Feng et al., 2018; Zurawski et al., 2017; Zurawski, Rodriguez, Hyde, Alford, & Hamm, 2016). 4.3 Materials and Methods 4.3.1 Tissue preparation and solutions All animal procedures complied with the National Institutes of Health of the USA 198! guidelines on animal care and were approved by Michigan State University Institutional Animal Use and Care Committee. Animal used for this chapter were between 5 to 10 weeks old. Only male animals were used due to the sex difference we have observed in our previous study (Chapter 3) (Feng et al., 2019). Mice were sacrificed by direct cervical dislocation and cerebellums were dissected and quickly mounted on a VibrotomeTM 1000 machine (Leica, Wetzlar, Germany). Sagittal cerebellar slices (250 µm) were prepared according to methods previously described by Yuan Y et al (Yuan & Atchison, 1999, 2003, 2007, 2016). Briefly, the cerebellums were transferred into a chilled oxygenated sucrose-based slicing solution and parasagittal cerebellar slices (250 µm thick) were cut using the VibrotomeTM 1000 machine (Leica, Wetzlar, Germany). The slicing solution contains (in mM): 125, NaCl; 2.5, KCl; 4, MgCl2; 1.25, KH2PO4; 26, NaHCO3; 0.5, CaCl2 and 25, D-glucose (pH 7.35–7.4 when saturated with 95% O2 /5% CO2 at room temperature of 22–25°C). Slices were incubated in the pre-chilled and oxygenated slicing solution for 15 min, and then transferred into standard artificial cerebrospinal fluid (ACSF) solution at room temperature for 30 min. The standard ACSF contains: 125, NaCl; 2.5, KCl; 1, MgCl2; 1.25, KH2PO4; 26, NaHCO3; 2, CaCl2 and 20, D-glucose (pH 7.35–7.4 saturated with 95% O2/5% CO2 at room temperature). 4.3.2 Electrophysiological recording Whole-cell patch clamp recording methods were detailed in previous publications (Yuan & Atchison, 1999, 2003, 2007, 2016). Slices were placed in a recording chamber 199! and perfused with standard ACSF bubbled with 95% O2/5% CO2. Individual neurons were visualized with a Nomarski 40X water immersion lens with infrared differential interference contrast optics using a Nikon E600FN upright microscope. Recording electrodes were fire polished and had a resistance of 3–7(MΩ when filled with pipette solution. For recording sIPSCs and mIPSCs, the pipette solution consisted of (in mM) 140, CsCl; 0.4, GTP; 2, Mg-ATP; 0.5, CaCl2; 5, Phosphocreatine Na2; 5, EGTA-CsOH; 10, HEPES (pH 7.3 adjusted with CsOH). For recording sEPSCs and mEPSCs, the pipette solution consisted of (in mM) 140, K-Gluconate; 0.4, GTP; 2, Mg-ATP; 0.5, CaCl2; 5, Phosphocreatine Na2; 5, EGTA-CsOH; 10, HEPES (pH 7.3 adjusted with KOH). The holding potential was −70(mV for recording of both IPSCs and EPSCs. For recording inhibitory currents, 6-Cyano-7-nitroquinoxaline-2,3-dione (CNQX, 10(µM) and amino-5-phosphonopentanoic acid (APV, 100(µM) were added to the external solution to block glutamate receptor-mediated sEPSCs. For recordings of miniature IPSCs (mIPSCs), 0.5(µM tetrodotoxin (TTX) was added to the external solution in addition to CNQX and APV. For recording sEPSCs, bicuculline (10 µM) was added to the external solution to block GABAergic receptor-mediated sIPSCs. For recording of miniature EPSCs (mEPSCs), the external solution was supplemented with 0.5(µM TTX in addition to bicuculline. Whole cell currents were filtered at 2–5(kHz with an 8-pole low-pass Bessel filter and digitized at 10–20(kHz for later off-line analysis using the pClamp 9.0 200! program (Molecular Devices, Inc., Sunnyvale, CA). All experiments were carried out at room temperature of 22–25°C. 4.3.3 Pharmacology The following agents were used: CNQX disodium salts (Sigma-Aldrich, St. Louis, MO), DL-2-Amino-5-phos-phon-o-pent-anoic (APV) acid solid (Sigma-Aldrich, St. Louis, MO), tetrodotoxin (TTX) (Tocris, Bristol, UK), pertussis toxin (PTX) (List Biological Laboratories, Campbell, CA), baclofen (Sigma-Aldrich, St. Louis, MO), N-ethylmalaeimide (NEM) (Sigma-Aldrich, St. Louis, MO), UK14,304 (Sigma-Aldrich, St. Louis, MO), CGP36216 (hydrochloride) (Cayman Chemical, Ann Arbor, MI), BRL44408 (Sigma-Aldrich, St. Louis, MO), cadmium chloride (Sigma-Aldrich, St. Louis, MO). All drugs were made up as 1000 x concentrated stock solutions in distilled water, aliquoted and stored at ~20oC. Aliquots were thawed and dissolved in oxygenated ACSF immediately prior to use. 4.3.4 SDS Page and Western Blots Male mice (6-8 weeks old) were sacrificed and their brains were dissected into different regions and flash-frozen in liquid nitrogen. For Western Blot analysis, tissues were thawed on ice and homogenized for 5 min with 0.5 mm zirconium beads in a Bullet Blender (Next Advance; Troy, NY) in RIPA buffer (20mM Tris-HCl, pH7.4, 150mM NaCl, 1mM EDTA, 1mM β-glycerophospate, 1% Triton X-100 and 0.1% SDS) with a protease inhibitor cocktail (Roche/1 tablet in 10 mL RIPA). Homogenates were centrifuged for 5 201! min at 4°C at(13,000 G. Supernatants were collected and protein concentrations determined using the bicinchoninic acid method (BCA method; Pierce; Rockford, IL). Protein concentration was normalized for all tissues with RIPA buffer and 2x SDS sample buffer containing β-mercaptoethanol (Sigma-Aldrich, St. Louis, MO) was added. Thirty µg of protein was loaded onto a 12% Bis-Tris gel (homemade), and samples were separated for 1.5 hrs at 160V. Proteins were then transferred to an Immobilon-FL PVDF membrane (Millipore, Billerica, MA) on ice either for 2 h at 100 V, 400 mA or overnight at 30 V, 50 mA. Immediately after transfer, PDVF membranes were washed and blocked in Odyssey PBS blocking buffer (Li-Cor, Lincoln, NE) for 40 min at RT. The membranes were then incubated with anti-Gαo (rabbit; 1:1,000; sc-387; Santa Cruz biotechnologies, Santa Cruz, CA) or anti-Gβ (recommended for detection of Gβ1, Gβ2, Gβ3 and Gβ4; mouse; 1:1000; sc-378; Santa Cruz biotechnologies, Santa Cruz, CA) and anti-actin (goat; 1:1,000; sc-1615; Santa Cruz) antibodies diluted in Odyssey blocking buffer with 0.1% Tween-20 overnight at 4oC. Following four 5-min washes in phosphate-buffered saline with 0.1 % Tween-20 (PBS-T), the membrane was incubated for 1 hr at room temperature with secondary antibodies (1:10,000; IRDye® 800CW Donkey anti-rabbit; IRDye® 800CW Donkey anti-mouse; IRDye® 680RD Donkey anti-goat; LI-COR Biosciences) diluted in Odyssey blocking buffer with 0.1 % Tween-20. The membrane was subjected to four 5-min washes in PBS-T and a final rinse in PBS for 5 minutes. The membrane was kept in the dark and the infrared signals at 680 and 800nm were 202! detected with an Odyssey Fc image system (LI-COR Biosciences). The Gαo polyclonal antibody recognizes an epitope located between positions 90-140 of the Gαo protein (Santa Cruz, personal communication). 4.3.5 Statistical Analysis Electrophysiological data analysis was performed as described previously (Yuan & Atchison, 1999, 2003, 2007, 2016). The individual performing the analysis was blinded to the genotype of the sample until all results were recorded. In brief, spontaneous synaptic currents were first screened automatically using MiniAnalysis software (Synaptosoft Inc., Decatur, GA) with a set of pre-specified parameters. They were accepted or rejected manually with an event detection amplitude threshold at 5(pA for sIPSCs/mIPSCs and 3pA for sEPSCs/mEPSCs as well as the kinetic properties (fast rising phase and slow decay phase) of the spontaneous events. Unless otherwise specified, synaptic events per cell collected over a 2-min period were averaged to calculate the frequency and amplitude of spontaneous synaptic currents. Amplitudes of currents were measured after subtraction of the baseline noise. MiniAnalysis-derived results were plotted in GraphPad Prism (GraphPad; LaJolla, CA). Results from more than one neuron from a single animal were averaged prior to statistical analysis. Some graphs, when indicated, do show points for each individual neuron while bar graphs and error bars are calculated from the per animal data. Data are presented as mean value ± SEM, where n=number of animals. 203! Statstical significance was determined using unpaired Student’s t-test unless stated otherwise. A p value < 0.05 was deemed as significant. Quantification of infrared (IR) Western blot signals was performed using Image Studio Lite (LI-COR Biosciences). Individual bands were normalized to the corresponding actin signals, and WT Gαo was set as control for each blot. All data were analyzed using GraphPad Prism 7.0 (GraphPad; LaJolla, CA). 4.4 Results Purkinje cells mediate the entire output of the cerebellar cortex; therefore any mechanisms able to modulate the firing pattern of Purkinje cells will influence cerebellar function. Purkinje cells fire spontaneously, even in the absence of glutamate input, and the pattern of firing is strongly influenced by GABAergic input. At least under experiment conditions, two types of inhibitory interneurons, the basket cells and the stellate cells largely convey this inhibitory input onto the Purkinje cells (Donato et al., 2008). Presynaptic neurotransmitter release is strongly regulated by G-protein-coupled receptors (GPCRs). Many GPCRs in the central nervous system are coupled to the Gαo protein, which belongs to the Gαi/o family. The activation of Gαo protein by GPCRs leads to the inhibition of voltage-gated calcium channels, inhibition of cAMP production, activation of G-protein coupled inward rectifying potassium channels (GIRKs) and also inhibition of synaptic vesicle release, all of which can be possible mechanisms of GPCRs-mediated inhibition of GABA release. 204! 4.4.1 Presynaptic GABA release is suppressed in the cerebellar Purkinje cells of Gnao1+/G203R mice Baseline recording of sIPSCs, which are due to AP-dependent GABA release, was isolated by adding 10 µM CNQX and 100 µM AP-V in standard ACSF. Interestingly, CNQX and AP-V significantly increased IPSC events in Purkinje cells (data not shown), which is consistent with previous observations (Brickley, Farrant, Swanson, & Cull-Candy, 2001). There is a significant decrease in sIPSC frequency in G203R mice comparing to that of their WT sibling (WT: 21.0 ± 1.7 Hz; G203R: 12.7 ± 2.5 Hz; Figure 4.1C & 4.1D), but no difference is detected in sIPSC amplitude (WT: 41.8 ± 8.5 pA; G203R: 36.7 ± 7.4 pA; Figure 4.1E & 4.1F). Data were recorded from 25 cells of 13 mice for WT and 21 cells from 9 mice for G203R. mIPSCs, which are due to AP-independent GABA release in the Purkinje cells were investigated by the additional application of 0.5 µM TTX. As reported previously in cerebellar Purkinje cells, TTX reduced mean IPSC frequency and amplitude to isolate mIPSCs (Bardo, Robertson, & Stephens, 2002; Boxall, 2000; Harvey & Stephens, 2004; Yuan & Atchison, 2003). Slices from G203R mice exhibited a marked reduction in mIPSC frequency compared to that of WT mice (Figure 4.1G - 4.1L). The effect on mIPSC frequency was greater than that on sIPSCs (75% vs 40% decrease). Data were recorded from 25 cells of 13 mice for WT and 21 cells from 9 mice for G203R. 205! (A, B) Representative Figure 4.1 Cerebellar Purkinje cells in brain slices from 4-6 week-old G203R mice display reduced GABAergic spontaneous synaptic currents (sIPSCs) and reduced miniature synaptic currents recording of spontaneous inhibitory postsynaptic currents in a cerebellar Purkinje cell from a 4 week-old mouse in the presence of 10 µM of CNQX and 100 µM of AP-V at a holding potential of -70 mV. (C, D) G203R mice showed a decrease in the frequency of sIPSCs. (E, F) No significant difference is observed in the amplitude of sIPSCs between WT and G203R mice. Unpaired Student’s t-test; **p=0.0086 WT (n=13 mice), G203R (n=9 mice). (G, H) Representative recording of spontaneous miniature inhibitory postsynaptic currents in a cerebellar Purkinje cell from a 4 week-old mouse in the presence of 10 µM (mIPSCs). 206! Figure 4.1 (cont’d) of CNQX, 100 µM of AP-V and 0.5 µM TTX at a holding potential of -70mV. (I, J) G203R mice showed a decrease in the frequency of mIPSCs. (K, L) No significant difference is observed in the amplitude of mIPSCs between WT and G203R mice. Unpaired Student’s t-test; **p=0.0011; WT (n=13 mice), G203R (n=9 mice). Recordings from each cell are shown as a data point but the bar graph, error bars, and statistical analysis was averaged data per animal. 4.4.2 Gαo blockers can reverse the enhanced inhibition of mIPSC frequency in G203R mice To investigate whether the reduced frequency of sIPSCs and mIPSCs is due to an enhanced signaling by the G203R mutant Gαo, we examined the effect of sulphydryl alkylating agent NEM, which uncouples pertussis toxin-sensitive Gαi/o subunits from receptors by modifying cysteine residues (Aktories, Schultz, & Jakobs, 1982). Similar to the effects of NEM on GABAergic mIPSCs in other brain slice preparations, 50 µM NEM significantly increased the frequency of mIPSCs in both WT and G203R cerebellar slices (WT: from 4.19 ± 0.57 Hz to 22.3 ± 0.4 Hz; G203R: from 1.52 ± 0.35 Hz to 20.8 ±1.4 Hz; Figure 4.2D & 4.2G) but did not affect the amplitude (WT: 11.1 ± 1.7 pA to 19.1 ± 3.1 pA; G203R: 16.0 ± 1.9 pA to 30.8 ± 5.5 pA; Figure 4.2F & 4.2H). Also, NEM eliminated the difference in the mIPSC frequency between WT and G203R mice (WT: 22.3 ± 0.4 Hz; G203R: 20.8 ±1.4 Hz; Figure 4.2D & 4.2F). Considering that NEM is not a selective Gαo protein blocker, we also examined the effects of PTX incubation on the AP-dependent and AP-independent IPSCs in cerebellar 207! Purkinje cells. PTX catalyzes the ADP-ribosylation of the α subunit of the heterotrimeric Gi/o family, thereby preventing the G proteins from interacting with GPCRs (Mangmool & Kurose, 2011). All slices in this study were subject to incubation in 1 µg/mL PTX for more than 6 hours before recording. To maintain comparable conditions to compare data with and without PTX, separate slices were used as control so that both groups underwent the 6-hour incubation. PTX incubation significantly increased the mIPSC frequency in slices from G203R mice but had no effects on WT mice (WT: 4.10 ± 0.70 Hz to 3.40 ± 0.68 Hz; G203R: 1.31 ± 0.16 Hz to 2.29 ± 0.30 Hz; Figure 4.3C & 4.3G). The amplitude of mIPSCs was not changed in either WT or G203R mice after PTX incubation (WT: 12.1 ± 1.6 pA to 9.7 ± 1.4 pA, G203R: 19.6 ± 1.6 pA to 15.8 ± 2.9 pA; Figure 4.3D & 4.3H). In contrast to mIPSCs, sIPSC frequency was not significantly affected by PTX in either WT or G203R mice (Frequency: WT: 13.8 ± 1.7 Hz to 10.9 ± 3.6 Hz; G203R: 9.1 ± 1.4 Hz to 12.5 ± 2.3 Hz; Amplitude: WT: 18.9 ± 3.4 pA to 19.2 ± 8.9 pA; G203R: 29.2 ± 4.1 pA to 24.4 ± 4.5 pA; Figure 4.3E-4.3F, 4.3I-4.3J). 208! Figure 4.2 The frequencies of mIPSCs were sensitive to NEM, an inhibitor of Gαi/o proteins. (A, B) Representative recordings showing mIPSCs traces in (A) WT and (B) G203R mice with the presence of 50 µM NEM. (C, D) NEM eliminated the difference in mIPSC frequency between WT and G203R. (E, F) No significant difference was observed in amplitude of mIPSCs after adding NEM. (G) NEM significantly increased the frequency of mIPSCs, (H) but with minor influence in the amplitude of mIPSCs. Unpaired Student’s t-test; WT (n=8 mice), G203R (n=7 mice). 209! Figure 4.3 A selective inhibitor of Gi/o, pertussis toxin (PTX), increased the frequency of mIPSCs in G203R but not WT mice. Slices were incubated in 1 µM/ml of PTX for >6 hrs pre-recording. Representative traces showed the example recordings of (A) mIPSCs and (B) sIPSCs in WT and G203R mice before and after PTX incubation. (C, G) PTX incubation significantly relived the Go mediated inhibition of mIPSC frequency in G203R mice, but not WT mice. Unpaired Student’s t-test; WT (n=5 mice), G203R (N=6 mice); **p=0.006. (D, H) PTX did not change the mIPSC amplitude of either G203R or WT mice. Neither frequency (E, I) nor amplitude (F, J) of sIPSCs was affected by PTX incubation. Unpaired Student’s t-test; WT (n=5 mice), G203R (n=6 mice); *p=0.03 between WT vs. G203R mice without PTX incubation. Results between WT and G203R were not significant. 210! 4.4.3 Presynaptic glutamate release is not affected by the G203R mutation in Gαo protein Purkinje cells receive glutamatergic inputs at their extremely elaborated dendrites from parallel fibers at the molecular layer of the cerebellar cortex and send GABA outputs to the deep nuclei (Tian & Zhu, 2018). To investigate whether the GOF mutation G203R also affects excitatory inputs on Purkinje cells, we recorded sEPSCs and mEPSCs from Purkinje cells. The AP-dependent excitatory postsynaptic currents (sEPSCs) were isolated with 10 µM bicuculline and AP-independent mEPSCs were recorded with the addition of 0.5 µM TTX. Although the GOF Gαo protein significantly decreased sIPSCs and mIPSCs frequency, EPSCs are not seemingly affected by the G203R mutation in Gαo. Neither frequency nor amplitude of sEPSCs and mEPSCs showed significant differences between WT and G203R slices (Figure 4.4; sEPSC frequency: WT: 1.50 ± 0.20 Hz vs. G203R: 1.36 ± 0.43 Hz; sEPSC amplitude: WT: 7.52 ± 0.92 pA vs. G203R: 6.42 ± 0.50 pA; mEPSC frequency: WT: 1.05 ± 0.16 Hz vs. G203R: 0.99 ± 0.46 Hz; mEPSC amplitude: WT: 6.39 ± 0.91 pA; G203R: 6.24 ± 0.60 pA). 211! Figure 4.4 G203R mutant slices show no difference in either spontaneous excitatory postsynaptic currents (sEPSCs) or miniature excitatory postsynaptic currents (mIPSCs). (A) sEPCSs were recorded from Purkinje cells at a holding potential of -70mV in the presence of 10 µM bicuculline. (B) 0.5 µM TTX was then added to the bath in order to record mEPSCs. (C, D & G, H) No significant difference in either frequency or amplitude was observed between WT and G203R. (E, F & I, J) No significant difference between WT and G203R was observed in mEPSCs either. Unpaired Student’s t-test; WT (n=5), G203R (n=5). 212! 4.4.4 Effects of G-protein coupled GABAB receptors on AP-independent GABA release onto Purkinje cells The effects of baclofen on AP-independent mIPSCs, isolated by the application of 0.5 µM TTX, were investigated. After recording baseline mIPSCs, baclofen (10 µM) was applied to the bath. Baclofen caused a clear reduction in mean mIPSC frequency in slices from both WT (from 5.47 ± 0.80 Hz to 1.19 ± 0.25 Hz, 78% inhibition) and G203R mice (from 1.24 ± 0.20 Hz to 0.65 ± 0.11 Hz, 48% inhibition). Baclofen was typically applied for 4 to 8 min in this and subsequent experiments. After baclofen application, the difference in mIPSC frequency still remains between WT (1.19 ± 0.25 Hz) and G203R (0.65 ± 0.11 Hz) mice (Figure 4.5C & 4.5E). This suggests a role of GABAB receptors in regulating AP-independent GABA release. Mean mIPSC amplitude was unchanged by baclofen (Figure 4.5F; WT: 12.8 ± 2.3 pA to 13.6 ± 4.3 pA; G203R: 17.8 ± 1.9 pA to 17.3 ± 1.6 pA) (Figure 4.5D). Interestingly, the application baclofen did not affect either frequency or amplitude of sIPSCs (Figure S4.2), suggesting that GABAB receptors do not regulate the AP-dependent inhibitory neurotransmitter release. To confirm that Gαo causes the baclofen-induced inhibition of mIPSC frequency as previously reported (Harvey & Stephens, 2004), PTX was used to block Gαo protein. The Gαo antagonist, 1 µg/mL PTX eliminated baclofen-induced inhibition of mIPSC frequency (Figure 4.5I & 4.5K & 4.5M), while exhibiting no effects on mIPSC amplitude (Figure 4.5J, 4.5L, 4.5N). PTX increased mean mIPSC frequency from 1.31 ± 0.16 Hz to 2.29 ± 0.30 213! Hz in G203R mice but did not change that of the WT mice (non-PTX: 4.10 ± 0.70 Hz; PTX: 3.40 ± 0.68 Hz). These data are consistent with a presynaptic role of Gi/o subunit in baclofen-induced inhibition of AP-independent GABA release onto Purkinje cells. Figure 4.5 Activating GABAB receptor with baclofen reduces mIPSC frequency but not amplitude. PTX incubation eliminates baclofen-induced inhibition of mIPSC frequency in WT and G203R mice. Representative traces showing the reduced mIPSC responses before and after adding baclofen (10 µM) in WT (A) and G203R (B) mice w/o PTX or baclofen. (C, E, G) Baclofen significantly decreased the frequency of mIPSC and the difference between WT and G203R remains though baclofen was present. Unpaired Student’s t-test; WT (n=6), G203R (n=6); ****p<0.001, ***p=0.003, *p=0.029 (WT vs. G203R), *p=0.014 (G203R w/o baclofen). (D, F, H) Amplitude remained unchanged regardless of the existence of baclofen. No significant change was observed in the (I, K, 214! Figure 4.5 (cont’d) M) frequency or the (J, L, N) amplitude of mIPSCs after adding baclofen in both WT and G203R mice. Unpaired Student’s t-test; WT (n=5), G203R (n=6). 4.4.5 Effects of G-protein coupled α2A adrenergic receptors on AP-dependent GABA release onto Purkinje cells Adrenoceptors are divided into three subtypes, α1, α2, and β receptors, which are coupled to Gq/11-, Gi/o- and Gs-proteins respectively (Kobilka et al., 1987; O'Rourke, Iversen, Lomasney, & Bylund, 1994). Previous publications reported the dual regulation of AP-dependent GABA release modulated by both Gi/o-coupled α2 receptors and Gq-coupled α1 receptors (Hirono & Obata, 2006). Here, we investigated whether Gi/o-coupled α2 receptors plays a role in regulating AP-dependent GABA release. We used a selective α2 receptor agonist UK14,304 to confirm whether α2 receptors drives the decrease in sIPSC frequency and whether G203R mutant enhanced this reduction in sIPSC frequency. We applied 10 µM UK14,304 in bath perfusion. UK14,304 greatly inhibited AP-dependent GABAergic IPSC (sIPSC) frequencies of both WT (from 10.7 ± 2.6 Hz to 4.48 ± 1.65 Hz) and G203R (from 3.38 ± 0.99 Hz to 1.03 ± 0.14 Hz) mice (Figure 4.6C & 4.6E & 4.6G). Interestingly, a significant difference remained between WT and G203R mice in mIPSC frequency after the application of UK14,304 (Figure 4.6E; WT: 4.48 ± 1.65 Hz vs. G203R: 1.03 ± 0.14 Hz). Like baclofen, the amplitude of mIPSC was not affected by the application of UK14, 304 (Figure 4.6D & 4.6F &4.6H). 215! Figure 4.6 The frequency of sIPSCs is modulated by α2AR receptors. Representative recordings showing the sIPSCs of (A) WT and (B) G203R mice before and after adding the selective α2AR receptor agonist UK14,304 (10 µM). (C, E, G) UK14, 304 significantly reduced the frequency of sIPSCs in both WT and G203R mice, and the frequency of sIPSCs remained lower in G203R mice after UK14,304 treatment comparing to WT. (D, F, H) The amplitude of sIPSCs was unaffected with UK14, 304 treatment. Unpaired Student’s t test; WT (n=8 mice), G203R (n=9 mice); *p<0.05 216! 4.4.6 α2A adrenergic receptor-induced inhibition of sIPSC frequency depends on activation of voltage-gated calcium channels It has been established that Gβγ subunits can act on multiple types of voltage-gated calcium channels to inhibit calcium influx from the extracellular space and decrease neurotransmitter release (Currie, 2010; Zamponi & Currie, 2013). For AP-dependent GABA release, high voltage activated calcium channels (N-type and P/Q-type calcium channels) are first activated by depolarization from AP-stimulated influx of sodium ions. Following the activation of calcium channels, GPCRs stimulate Go, which releases the Gβγ subunit. Since the neurotransmitter release is directly proportional to the extent of calcium influx from the extracellular space and the resultant changes in the intra-terminal calcium concentration (Wu & Saggau, 1997), we examined whether the UK14,304-induced inhibition of GABAergic IPSCs is dependent on the extracellular calcium concentration. To verify effects of low extracellular calcium on the UK14,304-induced inhibition of GABAergic IPSCs, we recorded baseline sIPSCs and UK14,304-induced inhibition of sIPSCs in the presence of 100 µM cadmium chloride (CdCl2). Cd2+ greatly decreased sIPSC frequency in both WT (from 17.3 ± 1.6 Hz to 1.01 ± 0.22 Hz) and G203R (from 10.9 ± 2.3 Hz to 1.46 ± 0.27 Hz) mice (Figure 4.7Aa-b, 4.7Ba-b & 4.7C & 4.7E). It also eliminated the inhibition of sIPSC frequency induced by application of 10 µM UK14,304 (Figure 4.7Ab-c, 4.7Bb-c & 4.7C & 4.7E). The amplitude of sIPSCs showed a trend toward a decrease with application of Cd2+ but the change 217! was not significant (Figure 4.7D & 4.7F). Interestingly, UK14,304 did not affect the frequency of mIPSCs in either WT or G203R mice (Figure S4.2). However, since mIPSCs do not depend on the activation of extracellular calcium channels, Cd2+ does not affect the frequency and amplitudes of mIPSCs in either WT or G203R mice (Figure 4.8). Moreover, baclofen induced decrease of mIPSC frequency was not affected by the inhibition of membrane calcium channels either (Figure 4.8). Although the difference between mIPSC frequency of WT and G203R is not significant in this figure, it is understandable since there is fewer n numbers for this experiment shown in Figure 4.8 comparing to experiment in Figure 4.5. 218! Figure 4.7 Cadmium-block of extracellular calcium influx suppresses the frequency of sIPSCs in both WT and G203R mice. Representative recordings showing the sIPSCs of (A) WT and (B) G203R in (a) ASCF with 100 µM AP-V and 10 µM CNQX, (b) Cd2+(100 µM)-ASCF with AP-V and CNQX, (c) Cd2+ (100 µM)-ASCF with AP-V, CNQX and 10 µM UK14,304. (C, E) 100 µM Cd2+ significantly reduced sIPSC frequency in both WT and G203R mice and blocks inhibition of sIPSC frequency induced by UK14, 304. (D, F) 100 µM Cd2+ did not affect amplitudes of sIPSCs. Unpaired Student’s t-test; WT (n=5 mice), G203R (n=5 mice); ****p<0.0001, ***p<0.001. 219! Figure 4.8 Cadmium-block of extracellular calcium influx does not affect the frequency and the amplitude of mIPSCs in both WT and G203R mice. Representative recordings showing the sIPSCs of (A) WT and (B) G203R in (a) ASCF with 100 µM AP-V, 10 µM CNQX and 0.5 µM TTX; (b) Cd2+(100 µM)-ASCF with AP-V, CNQX and TTX; (c) Cd2+ (100 µM)-ASCF with AP-V, CNQX, TTX and 10 µM baclofen. (C, E) 100 µM Cd2+ did not reduce the frequency of mIPSCs in either WT or G203R mice. Baclofen (10 µM) reduced the frequency of mIPSC with the presence of 100 µM Cd2+. (D, F) Similarly, 100 µM Cd2+ does not affect amplitudes of sIPSCs. Unpaired Student’s t-test; WT (n=5 mice), G203R (n=5 mice); *p<0.05. 220! 4.4.7 G203R mice exhibit decreased Gαo protein expression but no change in Gβ levels Gαo G203R mutation led to a reduction in Gαo protein expression in transiently transfected HEK293T cells (Feng et al., 2017). To see if this still stands in vivo, we tested Gαo protein expression in whole brain and also selected brain regions WT and G203R mice. Results showed a significant reduction in Go protein expression (50% of WT) in the whole brain (Figure 4.9A, 4.9C). The decrease in Go protein expression was significant in cerebellum, cortex, hippocampus and striatum of the G203R mice (Figure 4.9B, 4.9D). No significant change was observed in the brain stem and the olfactory bulb between the WT and G203R mice. The Gβγ subunits not only support the role of GPCR-Gα interaction, but also act directly and independently to regulate downstream signaling. The number of identified effectors of Gβγ has grown in recently years. They include some important targets like the GIRK channel, P/Q and N-type calcium channels, and the SNARE protein complex (Blackmer et al., 2005; Herlitze et al., 1996; Qin, Platano, Olcese, Stefani, & Birnbaumer, 1997; Wells et al., 2012; Zhang et al., 2004; Zurawski et al., 2017). The current model of heterotrimeric G protein function hypothesizes that the conformational changes in Gα lead to its dissociation from Gβγ to expose effector interaction surfaces on Gβγ. In fact, many of the GPCR-dependent physiological process inhibited by PTX (Go/i family) are mediated by the Gβγ subunits rather than the Gα subunits (Ikeda, 1996; Logothetis, 221! Kurachi, Galper, Neer, & Clapham, 1987; Stephens et al., 1994; Welch et al., 2002). Thus, the gain-of-function mutation G203R in Gαo may function though Gβγ subunits to regulate the neurotransmitter release. The expression level of Gβ did not change in G203R mutant mice (Figure 4.10). However, G203R is a GOF mutation with a decreased Gαo protein expression level and a normal Gβ protein level. This could lead to a constitutive increase in Gβγ protein function in G203R mice. Figure 4.9 G203R mice showed a significant decrease in Gαo protein expression. (A, C) Whole brain Gαo expression level decreased to about 50% in G203R mice’s brain lysates. (B, D) This reduction in expression was most significant in specific regions like cerebellum (CERE), cortex (CTX), hippocampus (HIP) and striatum (STR), while remained unaffected in brain stem (BS) and olfactory bulb (OB). All expression levels were normalized to that of WT accordingly. Unpaired Student’s t-test; WT (n=8), G203R (n=8); ****p<0.0001, *p<0.05. 222! Figure 4.10 G203R mice did not show any significant changes in Gβ expression in the brain. (A) A representative gel shows the Gβ protein expression patterns in each individual brain region, including olfactory bulb (OB), brain stem (BS), striatum (STR), hippocampus (HIP), cerebellum (CERE) and cortex (CTX). (B) Quantification of the protein expression levels is unchanged in G203R mice brain lysates comparing to those of their WT siblings. Unpaired Student’s t-test; WT (n=4), G203R (n=4). 4.5 Discussion G203R was deemed a gain-of-function mutation in our previous report, where it showed an enhanced ability to support α2A adrenergic receptor mediated inhibition of cAMP production (Feng et al., 2017). Patients with G203R all develop severe movement disorders and seizures at an early age (Arya et al., 2017; Kelly et al., 2019; Nakamura et al., 2013; Saitsu et al., 2016; Schirinzi et al., 2019; Schorling et al., 2017; Xiong et al., 2018). Previously, we reported that G203R mice, especially male mice, exhibit abnormalities in a battery of motor and behavioral tests (Feng et al., 2019). In this 223! chapter, we explore possible physiological mechanisms of GNAO1-related movement disorders. Purkinje cells function as the sole output of cerebellar neural signaling transduction. They synapse onto the deep nuclei of the cerebellum and release inhibitory neurotransmitter to control their interaction with the thalamus. Therefore, the altered excitatory/inhibitory regulation received by Purkinje cells can reflect possible abnormalities in the cerebellum. The cerebellum is well-known for playing a role in motor coordination and control of movement. Recent research has also linked disturbed cerebellar function to ataxia and dystonia (Bologna & Berardelli, 2018; Garcia et al., 2017; Marsden, 2018). Consequently, I started my investigation by examining neural control of cerebellar Purkinje cells. I discovered that the Gnao1 G203R mutation decreased both AP-dependent sIPSC frequency and AP-independent mIPSC frequency. Amplitudes were unaffected. Inhibition of Go signaling with NEM or PTX increased the frequency of mIPSCs. This finding suggests that the G203R mutation enhanced inhibition of GABA release through a pre-synaptic mechanism (Figure 4.11). Since PTX only increased the IPSC frequency in G203R but not WT slices, this suggests that G203R is a bona fide gain-of-function mutation. As such, the mutant Gαo protein has enhanced inhibition compared to a normal-functioning WT protein. Moreover, we have also confirmed that mIPSCs and sIPSCs are likely mediated by different receptors through different mechanisms. 224! AP-independent mIPSCs are mainly mediated by GABAB receptors (Figure 4.11C & 4.11D). However, GABAB receptors likely modulate spontaneous GABA release by inhibition of synaptic vesicle fusion through actions of Gβγ rather than inhibition of membrane calcium channels (Figure 4.11C & 4.11D). In contrast, AP-dependent sIPSCs are regulated by α2A adrenergic receptors through inhibition of voltage-gated calcium channels (Figure 4.11A & 4.11B). The identification of the relevant GPCRs provides a possible direction for new drug discovery and drug repurposing. Antagonists with combined effects on GABAB receptors and α2A adrenergic receptors may be an effective strategy for suppressing GNAO1-associated movement disorders. Figure 4.11 Models of GABABR and α2AR mediated inhibition of GABA release. (A) α2AR agonist activates α2AR, which results in the separation of Gαo and Gβγ. Gβγ inhibits 225! Figure 4.11 (cont’d) calcium influx from membrane calcium channels activated by AP. (B) G203R mutant Gαo protein enhances the suppression of calcium influx, which lead to a reduction in GABA release. (C) Spontaneous GABA release without AP stimulation is regulated by the activation of GABABR, which inhibits synaptic vesicle fusion. (D) G203R mutant Gαo further inhibits the synaptic vesicle release. However, it is not clear whether the G203R mutant Go heterotrimer inhibits the neurotransmitter release via the mutated Gαo subunit or by the released free Gβγ protein. Activated Gαo protein can expose the surface on Gβγ to form a core site for effector binding and effector activation (Smrcka, 2008). The G203R mutant mice exhibited a reduced mIPSC frequency in cerebellar Purkinje cells in the absence of any external GPCR agonists. This suggests three possibilities. First, endogenous agonists in the cerebellar slices may activate Go-coupled GPCRs. Second, sufficient free Gβγ subunits may exist, due to the reduced amount of mutant Gαo, and may cause pre-synaptic inhibition of GABA release. This is plausible considering that G203R mutant mice did not show any reduction in Gβ protein (Figure 4.10) but had a significant decrease in Gαo protein expression (Figure 4.9). Third, the G203R mutation may consecutively activate Gαo signaling pathways. This could activate both Gαo and also release free Gβγ subunits to mediate Gβγ signaling. There has been reported GNAO1 mutation that is consecutively active: Q205L (Ram, Horvath, & Iyengar, 2000). Structurally, G203 and Q205 are located close together; therefore it is reasonable to suspect that G203R may be a constitutively active GNAO1 mutation that does not require the activation by a 226! GPCR. Moreover, the Gαo G203R mutant showed a more rapid GDP release than the WT - as measured by the binding of a fluorescent GTP analog (personal communication). That suggests that the GOF effect of G203R may provide free Gβγ subunits more quickly than the WT Gα. Figure 4.12 Gβγ may play a major role in the regulation of IPSCs. (A) Activation of GPCRs leads to the separation of Gα-GTP and Gβγ. They may carry on content dependent activation or inhibition of the downstream signaling targets. (B) G203R mutant Gαo protein may contribute to an enhanced function of Gβγ, which leads to an increased inhibition of Gβγ-mediated inhibition of N- and P/Q-type calcium channels, AC, and synaptic vesicle fusion. But G203R may tamper the signaling pathway mediated by Gαo-GTP, like Gαo-activated neurite outgrowth. 227! A confusing aspect of the GOF GNAO1 mutations in children and in our mouse models has been that these mutants causing the dual phenotypes of MD and epilepsy. One potential explanation for this could be that GNAO1 G203R mutant mice may have context-dependent GOF and LOF in for different signal outputs (Figure 4.12) or in different brain regions (Figure 4.13). Despite the preliminary biochemical data showing G203R mutant Gαo protein’s rapid GDP release, we also found that protein expression of the G203R mutant is significantly lower that normal in cerebellum, striatum, cortex, and hippocampus (Figure 4.9). If one signal (e.g. neural migration) was mediated by Gα and the other (e.g. VGCC inhibitor) was mediated by Gβγ, the signal mediated by Gαo could be reduced (i.e. Gαo mediated activation of neurite outgrowth), while there is more free Gβγ causing GOF for the inhibition of AC, VGCC or vesicle release (Figure 4.12). There is one precedent for this with a human Gαs mutant that causes increased signaling in the testis but reduced signaling in the pituitary (Turan & Bastepe, 2015). In a similar aspect, the reason why G203R and R209H mutant mice showed different behavioral results can be attributed to that the different Gnao1 mutations tilt the balance between excitatory and inhibitory effects in different brain regions (Figure 4.13). If the G203R GOF mutant has a stronger influence on neurotransmitter release in brain regions that are closely related to epileptogenesis (i.e. cortex and hippocampus), while the R209H mutant with NF behavior does not, then it is more likely for G203R mutant mice to develop a higher susceptibility to seizures. 228! More research needs to be done to assess whether the G203R mutation could lead to different signaling outcomes in different brain regions. Future directions should focus on identifying the link between the cell-types and signal outputs that can test this model we proposed. Figure 4.13 GNAO1 mutations may have region specific effects, which cause an imbalance between excitatory and inhibitory neurotransmitters. Under normal condition, Go mainly inhibits the inhibitory neurotransmitter release to keep a fine-tuned balance between inhibitory and excitatory effects. However, G203R mutant and R209H mutant may reduce the inhibitory effects, hence overexcite the brain. Considering the difference in the presence of human symptoms, it is likely that G203R and R209H mutants affects brain regions that control movements like cerebellum or basal ganglia, but G203R mutant can further affect hippocampus and cortex therefore lead to the onset of epilepsy in both animals and humans. 229! APPENDIX 230! APPENDIX SUPPLEMENTAL DATA Figure S4.1 Despite the hypothesis that α2A receptor antagonist yohimbine (10µM) could reverse the inhibition of sIPSC frequency induced by UK14, 304, the application of yohimbine further reduced the sIPSC frequency with the application of UK14, 304. Representative traces are shown here with WT (A) and G203R (B) mice in the recoding of baseline level sIPSCs (a), sIPSCs with the application of UK14, 304 (b), 231! Figure S4.1 (cont’d) and sIPSCs with the application of both UK14, 304 and yohimbine (c). (C, E) Although not significant, yohimbine seems further reduced the sIPSC frequency that has already been decreased by the application of UK14, 304. It is highly possible since yohimbine is not a highly selective α2A receptor antagonist. It also antagonizes multiple serotonin receptors (Papeschi, Sourkes, & Youdim, 1971; Winter & Rabin, 1992), which also plays a role in regulating cerebellar GABA release and development (Nichols, 2011; Oostland & van Hooft, 2013). (D, F) Neither UK14, 304 nor yohimbine has any effects on the amplitudes of sIPSCs in WT and G203R mice. Unpaired Student’s t-test; WT (n=5), G203R (n=5). Figure S4.2 Baclofen does not affect either frequency or amplitude of sIPSCs, and UK14,304 does not affect mIPSC frequency or amplitude. Recording of sIPSCs are not affected by baclofen (10 µM) in either frequency (A) or amplitude (B). (WT: n=2 mice; 232! Figure S4.2 (cont’d) G203R: n=2 mice). Amplitudes (C) or frequency (D) of mIPSCs are not decreased by UK14, 304 (10 µM) either. (WT: n=2 mice; G203R: n=2 mice). Figure S4.3 Heterozygous G184S (GOF) mice also showed a low Gαo protein expression level. In whole brain lysates, (A) representative gel and (B) quantification of the relative protein level both showed that G184S mice had a reduced Gαo protein expression level (80% expression) comparing to WT (100%) and heterozygous KO mice (50%). Similar to the protein expression pattern seen in G203R mice, the Gαo protein levels were different in different brain regions. Onw-way ANOVA; +/+ (n=9), +/- (n=9), +/G184S (n=9); ****p<0.001, **p<0.01. (C, D) Heterozygous KO mice exhibited a reduced Gαo protein level in all brain regions tested: hippocampus (HIP), cortex (CTX), striatum (STR) and cerebellum (CERE). However, heterozygous G184S mice showed selective Gαo protein reduction in hippocampus and cerebellum but not in cortex or striatum. Two-way ANOVA; +/+ (n=6), +/- (n=6), +/G184S (n=6); ***p<0.001, **p<0.01. 233! Figure S4.4 Female Gnao1+/G184S mice showed reduced sIPSC frequency in hippocampal pyramidal cells, cortical layer II/IV pyramidal cells but not in cerebellar Purkinje cells. (A-D) Frequency of sIPSCs was significantly different in Gnao1+/G184S mice in the hippocampal CA1 pyramidal neurons (B, D), but amplitude showed no difference between WT and G184S mice (A, C). Unpaired Student’s t-test; WT (n=6), G184S (n=9); **p<0.01. (E-H) The same trend of a reduced sIPSC frequency was also seen in the cortical layer II/IV pyramidal neurons in the G184S mice (E, G), and the amplitudes of those two groups were not significantly different, although a trend of decreased amplitude can be seen in the G184S mice (F, H). Unpaired t-test; WT (n=8), 234! Figure S4.4 (cont’d) G184S (n=13); **p<0.01. (I, J) There is no significant difference in either frequency or amplitude in sIPSCs between WT and G184S mice’s cerebellar Purkinje cell. Unpaired Student’s t-test; WT (n=9), G184S (n=9). Figure S4.5 Brain Gαo expression does not change in R209H mice’s brain lysates. (A) Representative gel showing the Gαo protein expression in different brain regions of both WT (+/+) and R209H (+/R209H) mice. (B) Quantification of Gαo protein expression shows that there is no significant difference in all 6 brain regions tested between WT and R209H mice. CERE: cerebellum; OB: olfactory bulb; STR: striatum; HIP: hippocampus; BS: brain stem; CTX: cortex; Unpaired Student’s t-test; WT (n=7), R209H (n=7). 235! REFERENCES 236! REFERENCES Aktories, K., Schultz, G., & Jakobs, K. H. (1982). Inactivation of the guanine nucleotide regulatory site mediating inhibition of the adenylate cyclase in hamster adipocytes. Naunyn Schmiedebergs Arch Pharmacol, 321(4), 247-252. Arya, R., Spaeth, C., Gilbert, D. L., Leach, J. L., & Holland, K. D. (2017). GNAO1-associated epileptic encephalopathy and movement disorders: c.607G>A variant represents a probable mutation hotspot with a distinct phenotype. Epileptic Disord, 19(1), 67-75. doi:10.1684/epd.2017.0888 Barbour, B. (1993). Synaptic currents evoked in Purkinje cells by stimulating individual granule cells. Neuron, 11(4), 759-769. Bardo, S., Robertson, B., & Stephens, G. J. (2002). Presynaptic internal Ca2+ stores contribute to inhibitory neurotransmitter release onto mouse cerebellar Purkinje cells. Br J Pharmacol, 137(4), 529-537. doi:10.1038/sj.bjp.0704901 Blackmer, T., Larsen, E. C., Bartleson, C., Kowalchyk, J. A., Yoon, E. J., Preininger, A. M., . . . Martin, T. F. (2005). G protein betagamma directly regulates SNARE protein fusion machinery for secretory granule exocytosis. Nat Neurosci, 8(4), 421-425. doi:10.1038/nn1423 Bologna, M., & Berardelli, A. (2018). The cerebellum and dystonia. Handb Clin Neurol, 155, 259-272. doi:10.1016/B978-0-444-64189-2.00017-2 Boxall, A. R. (2000). GABAergic mIPSCs in rat cerebellar Purkinje cells are modulated by TrkB and mGluR1-mediated stimulation of Src. J Physiol, 524 Pt 3, 677-684. doi:10.1111/j.1469-7793.2000.00677.x Brickley, S. G., Farrant, M., Swanson, G. T., & Cull-Candy, S. G. (2001). CNQX increases GABA-mediated synaptic transmission in the cerebellum by an AMPA/kainate receptor-independent mechanism. Neuropharmacology, 41(6), 730-736. Brown, A. M., Arancillo, M., Lin, T., Catt, D. R., Zhou, J., Lackey, E. P., . . . Sillitoe, R. V. (2019). Molecular layer interneurons shape the spike activity of cerebellar 237! Purkinje cells. Sci Rep, 9(1), 1742. doi:10.1038/s41598-018-38264-1 Cesana, E., Pietrajtis, K., Bidoret, C., Isope, P., D'Angelo, E., Dieudonne, S., & Forti, L. (2013). Granule cell ascending axon excitatory synapses onto Golgi cells implement a potent feedback circuit in the cerebellar granular layer. J Neurosci, 33(30), 12430-12446. doi:10.1523/JNEUROSCI.4897-11.2013 Currie, K. P. (2010). G protein modulation of CaV2 voltage-gated calcium channels. Channels (Austin), 4(6), 497-509. doi:10.4161/chan.4.6.12871 Donato, R., Rodrigues, R. J., Takahashi, M., Tsai, M. C., Soto, D., Miyagi, K., . . . Edwards, F. A. (2008). GABA release by basket cells onto Purkinje cells, in rat cerebellar slices, is directly controlled by presynaptic purinergic receptors, modulating 521-532. influx. doi:10.1016/j.ceca.2008.03.006 Calcium, 44(6), Ca2+ Cell Eccles, J. C. (1967). Circuits in the cerebellar control of movement. Proc Natl Acad Sci U S A, 58(1), 336-343. doi:10.1073/pnas.58.1.336 Eccles, J. C., Llinas, R., & Sasaki, K. (1966a). The mossy fibre-granule cell relay of the cerebellum and its inhibitory control by Golgi cells. Exp Brain Res, 1(1), 82-101. Eccles, J. C., Llinas, R., & Sasaki, K. (1966b). Parallel fibre stimulation and the responses induced thereby in the Purkinje cells of the cerebellum. Exp Brain Res, 1(1), 17-39. Feng, H., Khalil, S., Neubig, R. R., & Sidiropoulos, C. (2018). A mechanistic review on Dis. movement Neurobiol disorder. GNAO1-associated doi:10.1016/j.nbd.2018.05.005 Feng, H., Larrivee, C. L., Demireva, E. Y., Xie, H., Leipprandt, J. R., & Neubig, R. R. (2019). Mouse models of GNAO1-associated movement disorder: Allele- and sex-specific differences in phenotypes. PLoS One, 14(1), e0211066. doi:10.1371/journal.pone.0211066 Feng, H., Sjogren, B., Karaj, B., Shaw, V., Gezer, A., & Neubig, R. R. (2017). Movement disorder in GNAO1 encephalopathy associated with gain-of-function mutations. Neurology, 89(8), 762-770. doi:10.1212/WNL.0000000000004262 238! Fremont, R., Tewari, A., Angueyra, C., & Khodakhah, K. (2017). A role for cerebellum in the hereditary dystonia DYT1. Elife, 6. doi:10.7554/eLife.22775 Garcia, A. M., Abrevaya, S., Kozono, G., Cordero, I. G., Cordoba, M., Kauffman, M. A., . . . Ibanez, A. (2017). The cerebellum and embodied semantics: evidence from a case of genetic ataxia due to STUB1 mutations. J Med Genet, 54(2), 114-124. doi:10.1136/jmedgenet-2016-104148 Gowen, E., & Miall, R. C. (2007). The cerebellum and motor dysfunction in 268-279. Cerebellum, 6(3), disorders. neuropsychiatric doi:10.1080/14734220601184821 Harvey, V. L., & Stephens, G. J. (2004). Mechanism of GABA receptor-mediated inhibition of spontaneous GABA release onto cerebellar Purkinje cells. Eur J Neurosci, 20(3), 684-700. doi:10.1111/j.1460-9568.2004.03505.x Herlitze, S., Garcia, D. E., Mackie, K., Hille, B., Scheuer, T., & Catterall, W. A. (1996). Modulation of Ca2+ channels by G-protein beta gamma subunits. Nature, 380(6571), 258-262. doi:10.1038/380258a0 Hirono, M., & Obata, K. (2006). Alpha-adrenoceptive dual modulation of inhibitory GABAergic inputs to Purkinje cells in the mouse cerebellum. J Neurophysiol, 95(2), 700-708. doi:10.1152/jn.00711.2005 Hull, C., & Regehr, W. G. (2012). Identification of an inhibitory circuit that regulates 149-158. Neuron, activity. 73(1), cell cerebellar doi:10.1016/j.neuron.2011.10.030 Golgi Ikeda, S. R. (1996). Voltage-dependent modulation of N-type calcium channels by 255-258. 380(6571), subunits. gamma Nature, G-protein doi:10.1038/380255a0 beta Jiang, M., & Bajpayee, N. S. (2009). Molecular mechanisms of go signaling. Neurosignals, 17(1), 23-41. doi:10.1159/000186688 Kelly, M., Park, M., Mihalek, I., Rochtus, A., Gramm, M., Perez-Palma, E., . . . Poduri, A. (2019). Spectrum of neurodevelopmental disease associated with the GNAO1 guanosine 406-418. doi:10.1111/epi.14653 triphosphate-binding Epilepsia, region. 60(3), 239! Kobilka, B. K., Matsui, H., Kobilka, T. S., Yang-Feng, T. L., Francke, U., Caron, M. G., . . . Regan, J. W. (1987). Cloning, sequencing, and expression of the gene coding for the human platelet alpha 2-adrenergic receptor. Science, 238(4827), 650-656. Konnerth, A., Llano, I., & Armstrong, C. M. (1990). Synaptic currents in cerebellar Purkinje cells. Proc Natl Acad Sci U S A, 87(7), 2662-2665. doi:10.1073/pnas.87.7.2662 Levitt, E. S., Purington, L. C., & Traynor, J. R. (2011). Gi/o-coupled receptors compete for signaling to adenylyl cyclase in SH-SY5Y cells and reduce opioid-mediated cAMP overshoot. Mol Pharmacol, 79(3), 461-471. doi:10.1124/mol.110.064816 Logothetis, D. E., Kurachi, Y., Galper, J., Neer, E. J., & Clapham, D. E. (1987). The beta gamma subunits of GTP-binding proteins activate the muscarinic K+ channel in heart. Nature, 325(6102), 321-326. doi:10.1038/325321a0 Mangmool, S., & Kurose, H. (2011). G(i/o) protein-dependent and -independent actions 884-899. (Basel), Toxins (PTX). 3(7), Pertussis of doi:10.3390/toxins3070884 Toxin Marsden, J. F. (2018). Cerebellar ataxia. Handb Clin Neurol, 159, 261-281. doi:10.1016/B978-0-444-63916-5.00017-3 Nakamura, K., Kodera, H., Akita, T., Shiina, M., Kato, M., Hoshino, H., . . . Saitsu, H. (2013). De Novo mutations in GNAO1, encoding a Galphao subunit of heterotrimeric G proteins, cause epileptic encephalopathy. Am J Hum Genet, 93(3), 496-505. doi:10.1016/j.ajhg.2013.07.014 Nichols, R. A. (2011). Serotonin, presynaptic 5-HT(3) receptors and synaptic plasticity in 5019-5020. J Physiol, 589(Pt 21), developing the cerebellum. doi:10.1113/jphysiol.2011.219782 O'Rourke, M. F., Iversen, L. J., Lomasney, J. W., & Bylund, D. B. (1994). Species orthologs of the alpha-2A adrenergic receptor: the pharmacological properties of the bovine and rat receptors differ from the human and porcine receptors. J Pharmacol Exp Ther, 271(2), 735-740. Oostland, M., & van Hooft, J. A. (2013). The role of serotonin in cerebellar development. Neuroscience, 248, 201-212. doi:10.1016/j.neuroscience.2013.05.029 240! Papeschi, R., Sourkes, T. L., & Youdim, M. B. (1971). The effect of yohimbine on brain serotonin metabolism, motor behavior and body temperature of the rat. Eur J Pharmacol, 15(3), 318-326. doi:10.1016/0014-2999(71)90098-7 Qin, N., Platano, D., Olcese, R., Stefani, E., & Birnbaumer, L. (1997). Direct interaction of gbetagamma with a C-terminal gbetagamma-binding domain of the Ca2+ channel alpha1 subunit is responsible for channel inhibition by G protein-coupled receptors. Proc Natl Acad Sci U S A, 8866-8871. doi:10.1073/pnas.94.16.8866 94(16), Ram, P. T., Horvath, C. M., & Iyengar, R. (2000). Stat3-mediated transformation of NIH-3T3 cells by the constitutively active Q205L Galphao protein. Science, 287(5450), 142-144. Reeber, S. L., Otis, T. S., & Sillitoe, R. V. (2013). New roles for the cerebellum in health and disease. Front Syst Neurosci, 7, 83. doi:10.3389/fnsys.2013.00083 Saitsu, H., Fukai, R., Ben-Zeev, B., Sakai, Y., Mimaki, M., Okamoto, N., . . . Matsumoto, N. (2016). Phenotypic spectrum of GNAO1 variants: epileptic encephalopathy to involuntary movements with severe developmental delay. Eur J Hum Genet, 24(1), 129-134. doi:10.1038/ejhg.2015.92 Schirinzi, T., Garone, G., Travaglini, L., Vasco, G., Galosi, S., Rios, L., . . . Leuzzi, V. (2019). Phenomenology and clinical course of movement disorder in GNAO1 variants: Results from an analytical review. Parkinsonism Relat Disord, 61, 19-25. doi:10.1016/j.parkreldis.2018.11.019 Schorling, D. C., Dietel, T., Evers, C., Hinderhofer, K., Korinthenberg, R., Ezzo, D., . . . Kirschner, J. (2017). Expanding Phenotype of De Novo Mutations in GNAO1: Four New Cases and Review of Literature. Neuropediatrics, 48(5), 371-377. doi:10.1055/s-0037-1603977 Smrcka, A. V. (2008). G protein betagamma subunits: central mediators of G protein-coupled receptor signaling. Cell Mol Life Sci, 65(14), 2191-2214. doi:10.1007/s00018-008-8006-5 Song, C. H., Bernhard, D., Hess, E. J., & Jinnah, H. A. (2014). Subtle microstructural changes of the cerebellum in a knock-in mouse model of DYT1 dystonia. Neurobiol Dis, 62, 372-380. doi:10.1016/j.nbd.2013.10.003 241! Stephens, L., Smrcka, A., Cooke, F. T., Jackson, T. R., Sternweis, P. C., & Hawkins, P. T. (1994). A novel phosphoinositide 3 kinase activity in myeloid-derived cells is activated by G protein beta gamma subunits. Cell, 77(1), 83-93. Sunahara, R. K., & Taussig, R. (2002). Isoforms of mammalian adenylyl cyclase: multiplicities of signaling. Mol Interv, 2(3), 168-184. doi:10.1124/mi.2.3.168 Tian, J., & Zhu, M. X. (2018). GABAB Receptors Augment TRPC3-Mediated Slow Excitatory Postsynaptic Current to Regulate Cerebellar Purkinje Neuron Response to Type-1 Metabotropic Glutamate Receptor Activation. Cells, 7(8). doi:10.3390/cells7080090 Turan, S., & Bastepe, M. (2015). GNAS Spectrum of Disorders. Curr Osteoporos Rep, 13(3), 146-158. doi:10.1007/s11914-015-0268-x Vanni, V., Puglisi, F., Bonsi, P., Ponterio, G., Maltese, M., Pisani, A., & Mandolesi, G. (2015). Cerebellar synaptogenesis is compromised in mouse models of DYT1 dystonia. Exp Neurol, 271, 457-467. doi:10.1016/j.expneurol.2015.07.005 Walker, R. H. (2016). The non-Huntington disease choreas: Five new things. Neurol Clin Pract, 6(2), 150-156. doi:10.1212/CPJ.0000000000000236 Welch, H. C., Coadwell, W. J., Ellson, C. D., Ferguson, G. J., Andrews, S. R., Erdjument-Bromage, H., . . . Stephens, L. R. (2002). P-Rex1, a PtdIns(3,4,5)P3- and Gbetagamma-regulated guanine-nucleotide exchange factor for Rac. Cell, 108(6), 809-821. Wells, C. A., Zurawski, Z., Betke, K. M., Yim, Y. Y., Hyde, K., Rodriguez, S., . . . Hamm, H. E. (2012). Gbetagamma inhibits exocytosis via interaction with critical residues on soluble N-ethylmaleimide-sensitive factor attachment protein-25. Mol Pharmacol, 82(6), 1136-1149. doi:10.1124/mol.112.080507 Winter, J. C., & Rabin, R. A. (1992). Yohimbine as a serotonergic agent: evidence from receptor binding and drug discrimination. J Pharmacol Exp Ther, 263(2), 682-689. Wu, L. G., & Saggau, P. (1997). Presynaptic inhibition of elicited neurotransmitter release. Trends Neurosci, 20(5), 204-212. Xiong, J., Peng, J., Duan, H. L., Chen, C., Wang, X. L., Chen, S. M., & Yin, F. (2018). 242! [Recurrent convulsion and pulmonary infection complicated by psychomotor retardation in an infant]. Zhongguo Dang Dai Er Ke Za Zhi, 20(2), 154-157. Yuan, Y., & Atchison, W. D. (1999). Comparative effects of methylmercury on parallel-fiber and climbing-fiber responses of rat cerebellar slices. J Pharmacol Exp Ther, 288(3), 1015-1025. Yuan, Y., & Atchison, W. D. (2003). Methylmercury differentially affects GABA(A) receptor-mediated spontaneous IPSCs in Purkinje and granule cells of rat cerebellar slices. J Physiol, 550(Pt 1), 191-204. doi:10.1113/jphysiol.2003.040543 Yuan, Y., & Atchison, W. D. (2007). Methylmercury-induced increase of intracellular Ca2+ increases spontaneous synaptic current frequency in rat cerebellar slices. Mol Pharmacol, 71(4), 1109-1121. doi:10.1124/mol.106.031286 Yuan, Y., & Atchison, W. D. (2016). Multiple Sources of Ca2+ Contribute to Methylmercury-Induced Increased Frequency of Spontaneous Inhibitory Synaptic Responses in Cerebellar Slices of Rat. Toxicol Sci, 150(1), 117-130. doi:10.1093/toxsci/kfv314 Zamponi, G. W., & Currie, K. P. (2013). Regulation of Ca(V)2 calcium channels by G receptors. Biochim Biophys Acta, 1828(7), 1629-1643. protein coupled doi:10.1016/j.bbamem.2012.10.004 Zhang, Q., Dickson, A., & Doupnik, C. A. (2004). Gbetagamma-activated inwardly rectifying K(+) (GIRK) channel activation kinetics via Galphai and Galphao-coupled receptors are determined by Galpha-specific interdomain interactions that affect GDP release rates. J Biol Chem, 279(28), 29787-29796. doi:10.1074/jbc.M403359200 Zurawski, Z., Page, B., Chicka, M. C., Brindley, R. L., Wells, C. A., Preininger, A. M., . . . Hamm, H. E. (2017). Gbetagamma directly modulates vesicle fusion by competing with synaptotagmin for binding to neuronal SNARE proteins embedded in 12165-12177. doi:10.1074/jbc.M116.773523 membranes. 292(29), J Biol Chem, Zurawski, Z., Rodriguez, S., Hyde, K., Alford, S., & Hamm, H. E. (2016). Gbetagamma Binds to the Extreme C Terminus of SNAP25 to Mediate the Action of Gi/o-Coupled G Protein-Coupled Receptors. Mol Pharmacol, 89(1), 75-83. doi:10.1124/mol.115.101600 243! CHAPTER 5: CONCLUSION AND FUTURE DIRECTIONS 244! 5.1 General conclusion It has been six years since the first reported cases of the GNAO1 mutation-associated neurological disorders in 2013 (Nakamura et al., 2013). There is a rapid increase in the number of GNAO1 mutations reported. Advances in neurological genetics and the growing interest in genetic counseling have pushed the increased interest in GNAO1 mutation-related movement disorders and/or epilepsy. Recently, The Bow Foundation (www.gnao1.org) has been founded in support of research for understanding GNAO1-associated neurological disorders. We are one of the first labs that took an interest in GNAO1-related neurological disorders. Through five years of research on my dissertation, I identified a genotype-phenotype correlation between the in vitro function of GNAO1 mutations and the nature of patients’ neurological symptoms (Chapter 2). Our lab has also established three animal models that phenocopy human GNAO1 patients and I have, in collaboration with Cassie Larrivee and Jeffrey Leipprandt, characterized their movement abnormalities and seizure propensity (Chapter 3). Furthermore, I have obtained electrophysiological data establishing altered cerebellar signaling in mice with the Gnao1 G203R mutation which causes a movement disorder and verified that G203R is a bona fide GOF mutation in the neural context (Chapter 4). Although I established that all functioning GNAO1 mutations (GOF and NF mutations) are associated with movement disorder patients, and all non-functioning mutations (LOF 245! and PLOF mutations) are related to epilepsy patients (Feng et al., 2017), this genotype-phenotype correlation was created based on a human engineered system with transiently transfected Gao mutants and the α2AR in HEK293T cells. To test this correlation in a physiological background, we have selected mutations that are either the most prevalent (GOF: G203R and NF: R209H) or related to the most severe epilepsy (LOF; ΔT191F197) to verify our genotype-phenotype correlation model. Interestingly, mice with the LOF mutation ΔT191F197 were abnormally small and developed severe behavioral seizure at around day 7 of life (P7). All mice with the ΔT191F197 mutation died before P16 and the strain was lost. Like human patients with G203R mutations, mice heterozygous for the G203R mutations (Gnao1+/G203R or G203R mice) behaved abnormally in our behavioral tests of movement and also showed heightened sensitivity to seizures, which was assessed by a PTZ kindling study (Feng et al., 2019). Comparably, the NF mutation R209H is only associated with movement disorders in both humans and mice (Larrivee et al., 2019); they do not show an epilepsy pattern. These behavioral tests established that our genotype-phenotype correlation stands in a physiological context across both mice and humans. However, there is an obvious sex difference in our animal models that has not been consistently observed in patients, perhaps due to the relatively small size of human GNAO1 patient population. Also in humans, the GNAO1 mutations could cause prenatal death of male embryos. G203R mice showed a male-dominant movement abnormality, while R209H mice have 246! symptoms that are equally severe in both male and female mice. Since our animal models exhibit symptoms similar to human GNAO1 patients, they make it possible to use those animal models to study the mechanisms of how GNAO1 mutations could lead to the onset of neurological disorders. Specifically, I used the patch clamp technique to measure both the excitatory and inhibitory neurotransmitter release in cerebellar slices of G203R mice and discovered that G203R mice exhibited decreased GABA release while glutamate release was unaffected. Also, it is possible to use the animal models to test new compounds or to repurpose drugs that are specifically effective for GNAO1 mutation-related disorders. While my work has covered preliminary aspects of GNAO1 mutation-associated neurological disorders, more research needs to be done to address unanswered questions that are beyond the scope of this dissertation. First, we have not tested the functions of mutant Gαo in any neuronal cell line or used different canonical pathways (such as Ca++ or K+ channel regulation) for characterization. Second, we have not explained why NF mutations also cause neurological disorders. It is possible that NF GNAO1 mutations could lead to other disturbed downstream signaling pathways but do not affect inhibition of cAMP. Third, since the majority of patients with GNAO1 mutations present with developmental delay and hypotonia at birth, it is yet to be established that GNAO1 mouse models (G203R and R209H) also present developmental issues. One other interesting question on development is how this disorder will progress when our 247! mutant mice become older. There haven’t been any reported human GNAO1 patients over 45 years old; therefore understanding the progression of this disorder could potentially prepare patients for any future complications. Last but not least, the sex difference in expression of neurological disorders, as mentioned before, needs to be verified as larger patient populations are reported. In our animal model, there is an obvious sex difference in the movement disorders or epilepsy but it differs among genotypes. In this chapter, I will discuss some major directions this project could take and provide some analysis for the development of each direction. Many of these ideas are based on preliminary data that I collected but did not have time to develop into a complete story. Those results are presented as an Appendix to this chapter 5.2 Testing the functional changes of a growing variety of GNAO1 mutations Our previous model of genotype-phenotype correlation was established on Gαo-mediated inhibition of cAMP production in HEK293 cells. However, the Gαo protein, either functioning by itself or through interaction with the Gβγ protein, regulates multiple essential intracellular effectors in its functional signaling pathways. Therefore, cAMP cannot be the sole evaluation upon which the GNAO1 mutations are examined. 5.2.1 Do GNAO1 mutations affect Go’s inhibition of high-voltage gated calcium channels (N- type & P/Q- type calcium channels)? Go’s inhibition of Ca2+ channels has received extensive scrutiny. Intracellular calcium 248! levels are important for neuronal signal transduction and neuronal development. Numerous hormones or neurotransmitters suppress Ca2+ channel currents (Dunlap & Fischbach, 1978). Later, Dunlap and colleagues showed that treating dorsal root ganglion with PTX blocks noradrenaline and GABA-mediated inhibition of Ca2+ channels (Holz, Rane, & Dunlap, 1986). Moreover, GTPγS, a non-hydrolysable GTP analog that binds to and activates G proteins, irreversibly potentiates agonist-mediated inhibition of Ca2+ channels (Holz et al., 1986; Scott & Dolphin, 1986) while GDPβS, a stable form of GDP, blocks it. Specifically, opioids activate their receptors in dorsal root ganglion neurons to suppress N-type Ca2+ channels (Jiang et al., 1998). Neurons lacking Go protein lose the opioid inhibitory effect (Jiang et al., 1998). As mentioned in Chapter 1, LOF mutations in CACNA1A (which encodes the P/Q type Ca2+ channel subunit) and CACNA1B (which encodes the N-type Ca2+ subunit) and GNAO1 GOF mutations lead to similar neurological symptoms. There are several ways to test how mutations in GNAO1 affect Go’s inhibition of calcium channels. The most traditional way is to measure the Ca2+ currents with the patch clamp technique. This was also used in the first published GNAO1 mutation case report in 2013 (Nakamura et al., 2013). In preliminary data, I transfected a previously established HEK293 cell line stably expressing the three subunits of N-type Ca2+ channels (G1A1 cell line with α1B-1, α2Bδ, β1B subunits) with plasmids for the α2AR and WT Gαo. In this system, I found that norepinephrine could inhibit calcium currents (Figure 249! S5.1) (Bleakman et al., 1995; McCool, Pin, Brust, Harpold, & Lovinger, 1996). Using this approach, it should be possible to transiently transfect these cells with all of the GNAO1 mutants and test their ability to inhibit the calcium currents with the patch clamp technique. However, with the ever-growing number of the GNAO1 mutations, patch clamp methods may be a time-consuming procedure for this aim. An alternative to patch clamp studies is to use the high-throughput calcium mobilization assay. There have been multiple reports describing high-throughput assays with multiple available calcium dyes to screen for N-type calcium channel blockers (Lubin et al., 2006; Zamponi, Striessnig, Koschak, & Dolphin, 2015; Zhang, Kauffman, Yagel, & Codd, 2006). Using this strategy, we could use either the N-type channel expressing G1A1 cell line or a neuronal cell line to screen the GNAO1 mutants’ effects on calcium currents in a relatively short time. A preliminary study with the Fluo-4 NW dye and the Hamamatsu’s FDSS µCell imaging system in the MSU Assay Core confirmed that G1A1 cell line does express N-type calcium channels (Figure S5.2A). Also the SH-SY5Y human neurobastoma cell line is a good candidate for this screening (Figure S5.3B). Evaluation of whether GNAO1 mutations alter membrane calcium channel function is extremely important. The findings would not only broaden our understanding of the genotype-phenotype correlation of the GNAO1 mutation-related neurological disorders, but also could determine the most predictive functional assay for drug repurposing or development. 250! 5.2.2 Do GNAO1 mutations affect Go’s activation of G protein-regulated inward rectifying potassium (GIRK) channels? Potassium channels on the plasma membrane are another intracellular effector of Go-mediated signaling. Not only have multiple studies documented the importance of this pathway, mutations in potassium channels were also reported to cause movement disorders as discussed in Chapter 1 (Luscher & Slesinger, 2010). In hippocampal pyramidal cells, serotonin and the selective GABABR agoinst baclofen hyperpolarize cells by increasing K+ channel conductance. The serotonin and baclofen responses are also ablated in PTX-treated cells (Andrade, Malenka, & Nicoll, 1986). Addition of GDPβS reduces the cell’s response to serotonin and baclofen, while GTPγS mimics the action of serotonin and baclofen (Andrade et al., 1986). Moreover, both purified bovine brain Go proteins as well as a recombinant form of Gαo proteins activate K+ channels in membranes from hippocampal pyramidal cells (Jiang & Bajpayee, 2009; Peleg, Varon, Ivanina, Dessauer, & Dascal, 2002; VanDongen et al., 1988). However, studies later focused on the role of Gβγ complex’s ability to activate the K+ channels while the Gα subunit mostly modulates the channel kinetics (Corey & Clapham, 2001; Huang, Jan, & Jan, 1997; Lei et al., 2000; Logothetis, Kurachi, Galper, Neer, & Clapham, 1987; Reuveny et al., 1994). Mutations in the Gαo protein may affect its role as a chaperone to release the Gβγ complex. 251! Apart from traditionally used patch clamp techniques, using high-throughput screening with thallium ions as a surrogate for potassium ions was also developed for screening potassium ion channel blockers (Beacham, Blackmer, M, & Hanson, 2010). Cell lines stably expressing GIRK channels (Lei et al., 2000) or primary hippocampal neurons are both available and good candidates for screening the effects of GNAO1 mutation on Go-mediated activation of GIRK channels (for both Gαo and Gβγ mechanisms). 5.2.3 How do GNAO1 mutations affect G protein-regulated neurite outgrowth? The initial formation of neurites during neuronal differentiation is commonly referred to as “neurite outgrowth”. This is the beginning point for neurogenesis, which is a crucial but long and winding journey in the development process. The Gαo protein is not only the most abundant membrane protein in the mammalian central nervous system, but is also highly enriched in neuronal growth cones. Elucidation of the effects of GNAO1 mutations’ on regulation of neurite outgrowth can be an essential step towards understanding GNAO1 mutation-associated developmental delay. Recent studies showed that the Gαo protein might directly stimulate neurite outgrowth. First, both Gαo and one of its interactors GRIN (G protein-regulated inducer of neurite outgrowth) are largely enriched in the growth cones, and activation of both can induce neurite outgrowth (Chen, Gilman, & Kozasa, 1999; Hwangpo et al., 2012; Strittmatter, Fishman, & Zhu, 1994; Strittmatter, Valenzuela, Kennedy, Neer, & Fishman, 1990). Second, dopamine-activated D2 252! receptors, which couple to the Gαo protein, induce neurite outgrowth in cortical neurons (Reinoso, Undie, & Levitt, 1996). Also, activation of CB1 receptors leads to the activation of downstream signaling converging on STAT3, which induces neurite outgrowth in Neuro2A cells (He et al., 2005; Jordan et al., 2005). Additionally, collapse of growth cones, induced by contact between neurites and a variety of molecules, can be inhibited by pertussis toxin (Igarashi, Strittmatter, Vartanian, & Fishman, 1993). Moreover, Gβγ is also involved in regulating neurite outgrowth. Research showed that Nerve Growth Factor (NGF) promoted Gβγ’s interaction with microtubules and stimulated microtubule assembly (Sierra-Fonseca et al., 2014). Also, GRK2i, which sequesters Gβγ, inhibited neurite formation, disrupted microtubules and led to neurite damage, while the Gβγ activator mSIRK stimulated neurite outgrowth (Sierra-Fonseca et al., 2014). A neurite outgrowth assay has been performed with PC12 cells (Figure S5.3) (Strittmatter, Fishman, et al., 1994; Traina, Petrucci, Gargini, & Bagnoli, 1998), Neuro2A cells (Georganta, Tsoutsi, Gaitanou, & Georgoussi, 2013; He et al., 2005), SH-SY5Y cells (Figure S5.4) (Paik, Somvanshi, & Kumar, 2019) and with primary isolated neurons (Lotto, Upton, Price, & Gaspar, 1999; Reinoso et al., 1996). A high-throughput neurite outgrowth assay kit has also been developed to reduce the labor of neurite staining and counting (Yeyeodu, Witherspoon, Gilyazova, & Ibeanu, 2010). Whether mutations in GNAO1 would affect its role in neurite outgrowth remains unanswered. 253! Additionally, for neurite outgrowth, one inevitable question is how activation of the Gαo protein (Strittmatter, Fishman, et al., 1994) and increases in cAMP levels (Aglah, Gordon, & Posse de Chaves, 2008) both lead to neurite outgrowth. These two competitive pathways may take control during different stages during neurite extension. It would be interesting to see how GOF GNAO1 mutations, which lead to enhanced suppression of cAMP production, regulate neurite outgrowth in vitro. Table 5.1 Comparison of clinical patterns of G203R and R209H patients Case No. Sex Age of onset Seizures Hypotonia Developmental Delay Chorea/athetosis 1 F 7 mo + + + 2 F 7 d + + + Dystonia Severe EEG ++ ++ diffuse irregular spike-and- slow-wave complex at 5 yr ++ EEG Severe MRI slow-wave bursts, migrating focal epileptiform discharges ++ delayed myelination at 1 yr, 3 mo; reduced cerebral white matter, thin corpus callosum at 4 yr, 8 mo Nakamura et al 2013 progressive cerebral atrophy with delayed myelination at 14 mo Saitsu et al 2015 MRI Reference 3 F 9 d + + not + impressive movement disorder but characterized delta and theta activity and rare multi- regional, bi- hemispheric epileptic activity + mild atrophy Dietel et al 2016 4 M 1 mo + + + ++ G203R 5 F 3 mo + + + + 6 F birth + + + + ++ multifocal and diffuse discharges, along with generalized -onset seizures ++ multifocal sharp waves, left temporal seizure pattern + background slowing ++ 7 F + birth (deceased at 12 mo) + ++ hypsarrhythmia 2 M 18 mo + + + 3 M 2 y + + 9 M 12 d + + + + 1 M 1 y + + + 5 M 3 y + + + R209H 4 M 10 mo + + + + 6 F 7 M 8 F 6 mo 6 mo 6 mo + + + + + + + + + + 8 F birth + + + + + ++ multifocal paroxysmal activities in both temporal hemispheres + NA + normal no no irregularities other than irregularities other than diffuse slowing diffuse slowing NA NA ++ normal ++ NA normal progressive diffuse cerebral atrophy and volume loss atrophy, thin corpus in callosum (2 atrophy (10 mild mo) cerebellum Arya et al 2017 y) Schorling et al 2017 Schorling et al 2017 normal Xiong at al. 2018 thin corpus callosum Schirinzi et al 2018 hypomyelination and atrophy Schirinzi et al 2018 normal Menke et al 2016 normal Kulkarni et al 2015 normal Kulkarni et al 2015 normal Dhamija et al 2016 global atrophy at 15 yr Ananth et al 2016 13 mo: frontal lobe volume loss normal Kelly et al Blumkin et 2019 al 2018 frontal lobe volume loss (13 mo) Kelly et al. 2018 5.3 Comparison between R209H and G203R mouse models We have studied behavioral abnormalities using our mouse models with the R209H and G203R Gnao1 mutations (Chapter 3) and also explored electrophysiological characteristics of cerebellar Purkinje cells in the G203R mouse model (Chapter 4). Needless to say, the two models exhibit some differences and similarities that may help 254! us understand the differences between patients with NF GNAO1 mutations (i.e. R209) and patients with GOF mutations (i.e. G203R). A comparison between patients with R209H and G203R mutations is shown in Table 5.1. All patients present with developmental delay from an early age. However, the G203R patients all exhibited seizure episodes while R209H patients very seldom developed seizure events. In addition, G203R patients are more likely to develop severe brain malformations, which can be seen from their MRI results. It is hard to say, however, if those malformations were caused by or contributed to epileptogenesis in those patients. In this section, I will discuss some ideas for future directions generated from the similarities and differences between R209H and G203R human patients and animal models. 5.3.1 Do G203R and R209H mouse models exhibit delayed development? Among reported cases with GNAO1-associated neurological disorders, all G203R patients (Arya, Spaeth, Gilbert, Leach, & Holland, 2017; Dietel, 2016; Nakamura et al., 2013; Saitsu et al., 2016; Schorling et al., 2017; Xiong et al., 2018) and R209H patients (Ananth et al., 2016; Dhamija, 2016; Kulkarni, Tang, Bhardwaj, Bernes, & Grebe, 2016; Menke et al., 2016) exhibit developmental delay. Although we have assessed movement abnormalities of adult mice, we are still unclear about whether our mouse models replicate this seemingly universal symptom for human G203R and R209H patients. Heyser (Heyser, 2004) published a very detailed milestone assessment for rodents that can be adopted by this project. Understanding the role of Gαo role in neural development 255! is crucial for understanding GNAO1-associated disorders, since developmental delay seems to be unrelated to the genotype-phenotype correlation that we established between epilepsy and movement disorders. Patients with both GOF/NF and LOF GNAO1 mutations exhibit developmental delay. So far, apart from Gαo’s regulation of neurite outgrowth (Strittmatter, Fishman, et al., 1994) and growth cone collapse (Igarashi et al., 1993) in vitro, there has been little research done on the role of Gαo in mammalian neuronal development (Tanaka, Treloar, Kalb, Greer, & Strittmatter, 1999). In addition to confirming whether G203R and R209H mice have developmental delay, how Gαo might play a role in regulating neuronal development should also be addressed. At the in vitro level, primary neurons from brains of G203R and R209H mutant mice can be isolated for assessing neurite outgrowth, axonal elongation, and growth cone development. Mechanistically, one interesting question is how G203R and R209H mutations in Gαo affect its interaction with GAP-43. GAP-43 (also called neuromodulin or B57) is a “growth” or “plasticity” related presynaptic protein that plays a key role in modulating growth cone signal transduction (Strittmatter, Valenzuela, Vartanian, et al., 1991), axonal growth and guidance (Goslin, Schreyer, Skene, & Banker, 1988), and synapse formation (Holahan, 2017). Homozygous mice lacking GAP-43 die in the early postnatal period (Strittmatter, Fankhauser, Huang, Mashimo, & Fishman, 1995) and heterozygous GAP-43 deficient mice survived but suffered from neuronal developmental defects that last through adulthood (Latchney et al., 2014). The amino-terminal domain 256! of GAP-43 promotes release of GDP from and binding of GTP to Gαo (Strittmatter, Igarashi, & Fishman, 1994; Strittmatter et al., 1990; Strittmatter, Valenzuela, Sudo, Linder, & Fishman, 1991). Co-expression of Gαo and GAP-43 can also be seen throughout mouse embryo development stages (Schmidt, Zubiaur, Valenzuela, Neer, & Drager, 1994). GAP-43 increases the GTPγS binding activity of Gαo (Jiang & Bajpayee, 2009; Yang, Wan, Song, Wang, & Huang, 2009). It is possible that GNAO1 mutants affect the Gαo and GAP-43 interaction by changing the guanine nucleotide exchange rate. Another interactor of Gαo protein that may be of interest here is GRIN1 (G protein-regulated inducer of neurite outgrowth 1) encoded by the GPRIN1 gene. GRIN1 binds to both Gαi and Gαo protein through the GRIN1 carboxyl-terminal region (Chen et al., 1999). Co-expression of GRIN1 and constitutively active Gαo protein (Q205L) induces neurite extensions in Neuro2A cells through the activation of Cdc42 (Nakata & Kozasa, 2005). Like GAP-43, GRIN1 also co-localizes with Gαo protein expression at neuronal dendrites and axons in different regions of adult mouse brains (Masuho et al., 2008). There are other possibilities for how Gαo protein may play a role in mouse development. Understanding the mechanisms of GNAO1-related neurodevelopmental disorders might provide unique insights into mechanisms of GNAO1-related movement disorder and epilepsy after birth. 257! 5.3.2 How does the G203R mutation in Gao lead to epileptogenesis? One puzzle between the patients with R209H and G203R mutations is why R209H patients very seldom exhibit epilepsy, while all of the G203R patients present with both epilepsy and movement disorders (Table 5.1) (Feng, Khalil, Neubig, & Sidiropoulos, 2018). This difference was also confirmed in our animal models with the PTZ kindling study (Chapter 3) where male G203R mutants have enhanced kindling responses to PTZ. We did not observe spontaneous seizures by G203R mice but we have not done EEG recordings so spontaneous seizures are not entirely ruled out. Three of the Gnao1+/G203R mutant mice did die in adulthood (Figure 3.1C) - similar to the G184S mutant mice that do have rare spontaneous seizures (Kehrl et al., 2014). The mechanisms of epileptogenesis in the G203R mutant are unclear. Since Gαo and Gβγ are involved in multiple aspects of neurobiology, possible mechanisms of G203R mutation-induced epileptogenesis includes: 1) altered neurotransmitter release; 2) a loss of subset of neurons; 3) altered neurite density and/or synaptogenesis; 4) changed membrane properties (lower threshold for activation); 5) altered cell morphology; and/or 6) malformation of cortical/hippocampal development. Activation of Gαo and Gβγ is well-studied for regulation of neurotransmitter release pre-synaptically (Stephens, 2009) and membrane potential post-synaptically (Beckstead & Williams, 2007; Newberry & Nicoll, 1985). It would not be surprising if the G203R GOF mutant has a stronger influence on neurotransmitter release in brain regions that are 258! closely related to epileptogenesis, while the R209H mutant with NF behavior does not. In addition, due to the role of Gαo and Gβγ in regulating neuronal development, it is possible that the Gao protein with the G203R mutation leads to malformation during one or more neural developmental stages. This hypothesis can be tested by carefully monitoring the developmental states of G203R mice at the behavioral, morphological and cellular levels. Malformations of cerebral cortical development are common causes of neurodevelopmental delay and epilepsy (Barkovich, Guerrini, Kuzniecky, Jackson, & Dobyns, 2012). The alteration of one or several developmental steps, including proliferation of neural progenitors, migration of neuroblasts, layer organization, or neuronal maturation may all lead to cortical malformation (Pang, Atefy, & Sheen, 2008). Previously, I have stained and observed the cerebellum region of adult Gnao1+/G184S mice, which exhibited similar behavioral abnormalities with Gnao1+/G203R mice (Chapter 3). There were no major abnormalities in the morphology of the cerebellum in the G184S mice expect for a slight decrease in lobule number (Figure S5.7). No staining for the G203R mutant mouse brain was done for the purpose of observing the gross morphology. Activated Gαo and Gβγ both play a role in regulating these developmental steps. The literature shows that stimulation of Gαo inhibited neuronal migration of the EP cells, which are a set of ~300 gut neurons begin to express Gαo at the time coincident with their migration along the stereotyped pathways (Copenhaver & Taghert, 1989; Horgan & 259! Copenhaver, 1998; Horgan, Lagrange, & Copenhaver, 1994). Also, Gβ1 knockout mouse embryos developed neural tube defects, abnormal actin organization, and microcephaly and then died at P2 (Okae & Iwakura, 2010). In the G203R mice or human patients, the G203R mutant may trigger an abnormal formation at one or multiple embryonic stages in cortex or hippocampus, resulting in susceptibility to epilepsy. For example, mutant Gαo may cause the failure of GABAergic neurons to migrate toward the cortex thus altering the excitatory/inhibitory balance. This could result in network hyperactivity (Wonders & Anderson, 2006). Or mutant Gαo may cause abnormal SNAP-25 function (Zurawski, Rodriguez, Hyde, Alford, & Hamm, 2016), which could lead to derangements of synaptic transmission in the hippocampus. 5.3.3 How do G203R and R209H mice differ in movement disorder phenotypes? Although dystonia and chorea/athetosis are both seen in G203R and R209H patients, G203R mice and R209H mice exhibit a striking difference in their movement disorder phenotypes. G203R mice were less able to remain on the RotaRod, had decreased capability of lifting up heavy weights, and had more abnormal gait characteristics like human patients with G203R mutation. They did not exhibit abnormalities in the open field test. In contrast, R209H mice mainly exhibited increased locomotor activity in the open field arena, which has not been seen in any previous Gnao1 mouse model. While it is understandable that mouse models will not precisely reflect human symptoms, it is still an interesting question on how mutations in the same gene cause different movement 260! phenotypes in rodents. One hypothesis is that R209H and G203R mutations affect different brain regions. Gαo expression seems to be ubiquitous in mammalian brains, however, due to the complex and diverse neural networking system in different regions, it is possible that G203R and R209H mutations have effects distinct from each other. Gαo staining is enriched in the cerebral cortex (particularly the molecular layer), the neuropil of the hippocampal formation, and in the striatum, substantia nigra pars reticulata, and in the molecular layer of the cerebellum (Worley, Baraban, Van Dop, Neer, & Snyder, 1986). Coincidently, dystonia is commonly linked with injury to the basal ganglia, thalamus, brainstem, and the cerebellum. Chorea is associated with disorders of the cerebral cortex, basal ganglia, cerebellum and thalamus (Sanger et al., 2010). Both disorders are seemingly associated with brain areas with significant Gαo protein expression levels. In chapter 4, I showed that the G203R mutation causes a reduced frequency of GABA release in cerebellar slices from G203R mice. However, we have not yet tested changes due to the G203R mutation in other brain regions, so we are not sure whether signaling in the striatum is also affected. The brain regions affected by the R209H mutation are not yet known. To quickly locate the brain regions related to the movement phenotype, we can do regional injection of oxotremorine (Pelosi, Menardy, Popa, Girault, & Herve, 2017) or pertussis toxin (PTX) to help exclude the irrelevant brain regions. This will be a crucial finding to help guide follow-up research on more detailed mechanisms of 261! the movement disorder and to find more targeted therapeutic methods for GNAO1-associated movement disorders. 5.3.4 How is sex involved in abnormalities of the Gnao1 mutants? GNAO1-associated neurological disorders are more prevalent in females than males overall according to currently published reports (Chapter 1), but it may also only be due to relatively small patient numbers. Alternatively, a more severe male phenotype would cause a premature death of the fetus. However, in animal models we have observed striking sex-related phenotypes. In G184S GOF mouse models, female mice are more prone to both abnormal movements and seizures; while in G203R mouse models, male mice developed more a significant phenotype compared to female for both movement disorders and epilepsy. Interestingly, R209H mice do not seem to differ much in sex-related phenotypes. Although sex differences in human GNAO1 patients remains unclear, it is still interesting to consider the sex differences of the Gnao1 mutant animals in terms of their behavioral abnormalities. Biological differences between the male and female sexes contribute to many sex-specific illness and disorders. These differences are not only due to gonadal hormone secretion-related events, such as differences in neuroanatamy, synaptic patterns, and neuronal density, but also due to non-hormonal related aspects, particularly direct gene products mediated by genes located on the X- and Y- chromosomes (Ngun, Ghahramani, Sanchez, Bocklandt, & Vilain, 2011). 262! Most researchers attribute the sex differences in neurological disorders to the actions of estrogens, progestins, and androgens. After all, those hormones regulate early neurodevelopment to program the brain to be sexually bimorphic, and later activate circuitries that trigger adult behaviors after puberty (Kight & McCarthy, 2014). The actions of gonadal hormones lead to differences in brain structure (Farrell, Gruene, & Shansky, 2015; Phan et al., 2012), connectivity (Ingalhalikar et al., 2014), signaling (Harte-Hargrove, Varga-Wesson, Duffy, Milner, & Scharfman, 2015; Skucas et al., 2013), responsivity (Garrett & Wellman, 2009), plasticity (Gould, Woolley, Frankfurt, & McEwen, 1990; Greenough, Carter, Steerman, & DeVoogd, 1977; Parducz et al., 2006), and even adult neurogenesis (Galea, Spritzer, Barker, & Pawluski, 2006; Livneh & Mizrahi, 2011; Vivar, Peterson, & van Praag, 2016). This makes hormone levels and functions an important aspect to consider for the mechanisms of sex differences in our Gnao1 mutant mouse models. Gαo may directly play a role in this hormonal regulation. Estrogen attenuates the reuptake of both endogenous and exogenous dopamine in the striatum and nucleus accumbens by altering the D2 receptor responsiveness (Thompson & Certain, 2005). Estrogen also destabilizes GABAB, 5-HT1a, 5-HT1b and CB1 receptors after a short exposure (Mize & Alper, 2000). Progesterone activates progestin membrane receptors to down-regulate adenylyl cyclase activities, which can be blocked by PTX (Thomas et al., 2007). Nuclear localization of androgen receptors is also controlled by Gi-specific RGS proteins (Rimler, Jockers, Lupowitz, & Zisapel, 2007). 263! Emerging research has also shown that sex differences are mediated by mechanisms other than action of the hormone secretions. If hormones do not explain the sex differences, then we should consider the different effects of XX versus XY sex chromosome complement. After all, every neuron, glia, or other cell type carries either the male chromosomes (XY) or female chromosomes (XX), but not both (Arnold & Burgoyne, 2004). Although the traditional view attributes sex differences in neuronal development to different hormonal exposure, recent research shows that some non-gonadal tissues, including the brain, are sexually dimorphic even when they develop in a similar endocrine environment (Arnold & Burgoyne, 2004). For example, primary cell cultures harvested from the XX and XY mesencephalon and diencephalon before the differentiating actions of gonadal hormones are present can develop into different numbers of dopamine neurons or prolactin neurons, respectively (Beyer, Kolbinger, Froehlich, Pilgrim, & Reisert, 1992; Beyer, Pilgrim, & Reisert, 1991). Another example is the fact that male and female mammalian embryos develop at different rates at ages before the onset of gonadal differentiation. There is evidence supporting that genes on the Y chromosome enhance the rate of embryonic development (Burgoyne, 1993; Burgoyne et al., 1995); while X chromosome genes slow down development (Thornhill & Burgoyne, 1993). Another aspect is the presence of the SRY gene on the Y-chromosome. Research shows that animals with Sry gene are associated with a higher number of tyrosine hydroxylase (TH+) cells compared to those without Sry (Ngun 264! et al., 2011). In humans, SRY expression is seen in both adult and fetal brains (Clepet et al., 1993; Mayer, Lahr, Swaab, Pilgrim, & Reisert, 1998). Also Sry has a direct effect on the expression of TH in the substantia nigra in the rat (Dewing et al., 2006). Effects of the SRY gene are suspected to contribute to the susceptibility of men to Parkinson’s disease (PD). Previous research has not yet established the interconnection between functions of Gαo and other embryonic developmental events in mammals; therefore it is hard to determine whether Gαo plays a role in the embryonic development before or after gonadal hormones are present, or both. Besides, two different GOF mutations, G184S and G203R, seem to affect a different sex in terms of the severity of the motor behavior and seizures in mice (Chapter 3). This observation provides additional evidence of a pathophysiological mechanism more complex than just Gαo-regulated inhibition of cAMP. The sex differences in the G184S and G203R mice should each be established as a separate project with a deeper understanding of the differences between the G184S and G203R mutant mice. Here, I will mainly focus on the discussion of the male-dominant phenotype in the G203R mutant mice. From a hormonal secretion aspect, it is possible that estrogen exerts a neuroprotective effect in the female mice (Brann, Dhandapani, Wakade, Mahesh, & Khan, 2007; Green & Simpkins, 2000); therefore the male G203R mice have a more severe genotype. Research showed that estrogen reduced dopamine D2 receptors by 20% 265! to 25% (Chavez et al., 2010), which could counterbalance the GOF mechanisms induced by the G203R mutant in female mice, and leave the male mice affected. Another possibility would be that Gαo with the GOF G203R mutation affects the nuclear localization of androgen receptors in the hippocampus (Rimler et al., 2007), which reduces the serum androgen level, and consequently lead to reduction of the GABAergic inhibition (Frye, 2006; Reddy & Jian, 2010). This would expose the male G203R mice to the risk of disease such as temporal lobe epilepsy (Harden & MacLusky, 2004, 2005; Herzog, 1991). From a non-hormonal perspective, we should look into the embryonic developmental stages of the G203R mutant mice to see if it correlated with the Y-chromosome regulated increase in rate of development. Otherwise, we should also consider the SRY gene’s modulation of dopamine biosynthesis and motor function. SRY, as a transcription factor, can directly activate the TH promoter, which enhances expression of tyrosine hydroxylase, the rate-limiting enzyme of dopamine synthesis (Czech et al., 2012). Although it is yet to be established whether the G203R mutation could affect dopamine transmission or the survival of dopamine neurons in the brain, the presence of the SRY gene on the Y-chromosome in the male G203R mutant mice could potentially exacerbate the effect of increased dopamine in causing abnormal movements (Cepeda, Murphy, Parent, & Levine, 2014). 266! While we are not yet sure of the effect of the R209H mutation on Gαo’s functional pathways, R209H mutants may not exhibit as extensive an effect as G203R mutants do. This may be why we do not see a significant sex difference in the R209H mutant mice. 5.4 Development of a high-throughput assay for drug repurposing or drug development Calcium ions participate in a variety of physiological and pathological mechanisms. In presynaptic neurons, calcium levels regulate neurotransmitter release. Detectable calcium signaling results from a complicated interaction between activation and inactivation of both intracellular and extracellular calcium channels. This interaction can be observed with calcium-sensing probes such as Fura-2 or Fluo-4. Neural calcium signaling often results in sequential regenerative discharges of stored calcium, a process referred to as calcium oscillation (Dupont, Combettes, Bird, & Putney, 2011). Calcium oscillations can be categorized into two classes based on the involvement of intracellular calcium storage. The intracellular release of calcium most commonly derives from the endoplasmic reticulum (ER) driven by inositol 1,4,5-trisphosphate (InP3) (Streb, Irvine, Berridge, & Schulz, 1983). However, in excitable cells like neurons, the increase of intracellular calcium levels can also be initiated by activation of membrane channels that lead to the influx of calcium ions from the extracellular space (Tsien et al., 1986). This may further facilitate intracellular calcium release from the endoplasmic reticulum (ER) into the cytoplasm (Fabiato, 1983). 267! Dissociated mouse cortical neurons can re-associate in vitro to form connected synaptic networks. This method provides an assay to reduce the expense and labor of testing drugs in vivo and provides a potential approach for high-throughout screening. Previous research has shown that neural calcium oscillations involve the activation of NMDA, AMPA/kainate receptors, and mGluR (Dravid & Murray, 2004; Nash et al., 2002; Robinson, Kawahara, et al., 1993). Also, the rising phase of each calcium spike is usually coincident with a brief burst of action potentials (Murphy, Blatter, Wier, & Baraban, 1992). Additionally, calcium oscillations measured from dissociated rat cortical neurons are dependent on the influx of extracellular calcium rather than mobilization from intracellular stores (Wang & Gruenstein, 1997). Oscillating calcium activity in dissociated neurons is also temporally correlated with the maturity of neurons and the time in culture to allow connections between neurons. Generally, synchronized calcium oscillations occur in around DIV7-DIV10 after plating and they plateau at DIV13 (Pacico & Mingorance-Le Meur, 2014). This is coincident with a burst of synapse formation that happens around DIV14 (Ichikawa, Muramoto, Kobayashi, Kawahara, & Kuroda, 1993). Thus, the synchronized calcium-spiking events seem to be an emergent property due to the formation of a large number of glutamatergic synapses. In agreement with earlier studies (Robinson, Kawahara, et al., 1993; Robinson, Torimitsu, Jimbo, Kuroda, & Kawana, 1993; Wang & Gruenstein, 1997), my preliminary data showed that removal of extracellular Mg2+ induces cultured mouse neurons to undergo synchronized calcium 268! oscillations (Figure S5.5). Calcium oscillation activities are also dependent on extracellular calcium concentrations (Figure S5.5). Although previous studies showed that inhibition of voltage-gated K+ channels induces calcium oscillation (Wang & Gruenstein, 1997), we found that high levels of extracellular K+ also suppress calcium oscillation activity (Figure S5.5). Since the calcium oscillations shown here are uniform and stationary, the analysis can be accomplished by Fast Fourier Transform (FFT) that will identify both the frequency and the amplitude of these oscillations (Barhoumi, Qian, Burghardt, & Tiffany-Castiglioni, 2010; Vajda, Donnan, Phillips, & Bladin, 1981). Here, we intend to use this technique to explore whether there is any difference between WT and G203R mice in the calcium oscillatory activity. As discussed previously, the Gαo protein mainly functions to inhibit synaptic neurotransmitter release. However, since Gαo is present at both excitatory and inhibitory synapses, it is hard to predict the end result of abnormalities in Gαo regulation. Based on what we observed in Chapter 4, the G203R GOF Gαo protein mainly enhanced presynaptic inhibition of GABA release but has less to no inhibitory effect on glutamate release. Therefore, we hypothesize that the GOF G203R mutant enhances suppression of GABA release thereby increasing excitability of the neural network, which leads to increased calcium spiking events. Our preliminary study confirmed our hypothesis that G203R mutant animal cortical cell cultures (both Gnao1+/G203R and Gnao1G203R/G203R) have more calcium spikes (Figure S5.6). However, it has been challenging to optimize this technique for high-throughput 269! screening. The data generated between experiments are highly variable; therefore comparison between plates in multiple experiments done at different times requires a standard control for each plate. Another issue is the variability between animals, which is hard to avoid but can be minimized by increasing the n number for experiments. With further optimization, calcium oscillation can utilize primary neurons from different brain regions to address the different oscillatory activities between WT and mutant mice. Ultimately, this may be used in high throughput methods to identify candidate drugs or receptor targets for drug repurposing or new drug development. 5.5 Impact of work in this thesis on the field This thesis is the summary of the first project to explore the GNAO1-related rare but serious neurologic disorders in children. I defined the biochemical correlation between the functional changes of the human GNAO1 mutants and the patients’ neurological symptoms. In addition, our work also provided the first human avatar Gnao1 mutant mice and verified their value in studying the neurophysiological mechanisms of GNAO1 mutations. Subsequent mechanistic and intervention studies should greatly enhance the development of potential therapeutic strategies for these devastating childhood neurologic disorders. In a broader aspect, due to the ubiquitous expression of Gαo (encoded by GNAO1) and its multiple effectors, our study may enhance understanding of neurological disorders that involve the same pathways shared by the Gαo protein. Similar pharmacological interventions may also be valuable for genetic conditions 270! involving GNB1, ADCY5, PDE10A and GNAL. Furthermore, our study may serve as a prototype for other correlations between reported monogenic mutations and human neurological disorders. 271! APPENDIX 272! APPENDIX SUPPLEMENTAL DATA Figure S5.1 α2AR activates Gαo, which inhibits N-type calcium channels in G1A1 cells. GOF mutation G184S enhance the inhibition of calcium currents. (A) G1A1 cells stably express N-type calcium channels, which can be blocked by MVIIa but not entirely. Previous report showed G1A1 cells have endogenous L-type calcium channels. All membrane calcium channels can be blocked by Cd2+ ions and enhanced by Ba2+ ions. (B) G1A1 cells have negligible amount of Gαi/o proteins, therefore activating α2AR alone cannot inhibit the calcium currents. (C) G1A1 cells with transiently transfected Gαo 273! Figure S5.1 (cont’d) protein and α2AR can lead to the inhibition of calcium current when α2AR is activated by norepinephrine (NE). (D) GOF GNAO1 mutation G184S causes a slight enhanced inhibition of calcium current immediately after adding selective α2AR agonist UK14, 304, however, the maximum inhibition between WT and G184S is not significant. WT (n=6), G184S (n=7), pcDNA (n=3). (E) G184S (n=4) significantly enhanced inhibition of N-type calcium channels comparing to WT (n=4). Unpaired Student’s t-test; p=0.0431. Figure S5.2 Both G1A1 and SH-SY5Y cells are good candidates to study mutant Gαo’s effects on N-type calcium channels with Fluo 4-NW dye in a Hamamatsu µCELL plate reader. (A) N-type calcium channels in G1A1 cells are activated by 90 mM 274! Figure S5.2 (cont’d) KCl with the presence of 5 mM CaCl2. (B) SH-SY5Y cells have more types of endogenous calcium channels. All calcium currents can be induced by 90 mM KCl. N-type calcium channels can be blocked by 250 nM MVIIa, and L-type calcium channels can be inhibited by 10 µM Nifedipine. Figure S5.3 Neurite outgrowth in rat pheochromocytoma cells (PC12) can be induced by 50 ng/mL Nerve Growth Factor (NGF) in normal growth medium with reduced serum. (A) Representative pictures showing PC12 cells growth after 48 hrs in (a) normal growth medium, (b) normal growth medium with reduced FBS (3% FBS), and (c) 50 ng/mL NGF in normal growth medium with 3% FBS. (B) NGF significantly promotes the percent of cells with visible neurites (n=3; Unpaired Student’s t-test; p<0.001). Spontaneous neurite outgrowth does occur without any NGF. (C) NGF also significantly increases the neurite length from PC12 cells (n=3; Unpaired Student t-test; p<0.0001). 275! Figure S5.4 Neurite outgrowth can be induced by 10 µM retinoid acid (RA) in normal growth medium with reduced serum (30% FBS) in human neuroblastoma cells (SH-SY5Y). (A) SH-SY5Y cells are largely non-differentiated in normal growth medium with 10% FBS (a); reduced serum induces morphological changes and neurite outgrowth in SH-SY5Y cells (b); 10 µM RA in 3% FBS medium induces the majority of cells to differentiated (c). (B) Preliminary studies show that RA increases the percent of cells with neurites, (C) but does not seemingly affect the average neurite length. Since the n number is small, there is no significance in comparison but the trend of change is obvious. 276! Figure S5.5 Representative traces of modulations in calcium oscillations with different ion concentration in mixed cortical cultures from a WT mouse. (A) Extracellular calcium concentration is crucial for calcium spiking events. The frequency of calcium oscillation increases with lower calcium concentrations but the amplitudes of calcium spikes peaks with a calcium concentration between 6.3 mM and 3.1 mM. (B) Higher concentrations of magnesium inhibit calcium oscillatory activities and lower magnesium concentrations induce them. All magnesium concentration responses were done with 5 mM calcium present. (C) When potassium concentration is above 15 mM, calcium oscillatory activities are completely quiescent. Lower concentrations of potassium facilitate calcium oscillation. All potassium concentration responses were done with 5 mM calcium present. 277! Figure S5.6 High-throughput assessment of neural excitability in cortical cultures. Cortical cultures (neurons and glia) from P0 Gnao1+/G203R mutant mice and WT littermates (triplicate wells shown) were prepared in 96‐well plates, allowed to form 278! Figure S5.6 (cont’d) connections (14 DIV), labeled with Fluo-4-NW calcium indicator, then read in a Hamamatsu Cell reader for 10 minutes. Power spectra ‐ (A) (all) & (B) (mean ± SD) of 10 WT (96 wells), 10 heterozygous (Het 66 wells), and 2 homozygous (Homo 24 wells) pups under conditions showing differences in epilepsy studies (CaCl2 5 mM and KCl 7.5 mM), (C) AUC of Ca2+ power spectra for all wells. Due to the non‐ normal distribution, statistical analysis used the non‐parametric Kolgov‐Smirnov test of cumulative distributions (GraphPad Prism 8.1). (D) Heat map of Ca2+ levels ‐ 3 wells per mouse (4 WT & 2 Het) at CaCl2 5 mM and MgCl2 7.5 mM. Figure S5.7 Nissl staining compares gross morphological changes in cerebellum from Gnao1+/G184S and Gnao1+/+ mice at age 8 weeks old. (A) No obvious gross difference in cerebellum region of Gnao1+/G184S and Gnao1+/+ mice. (B) Lobule number is slightly lower in Gnao1+/G184S mice (4 slices were measured from each mice; Gnao1+/+: n=8, Gnao1+/G184S: n=7). (C) Molecular layer thickness of IV, V, VIa and VIb in the cerebellar region of Gnao1+/G184S and Gnao1+/+ mice does not show significant difference 279! Figure S5.7 (cont’d) (12 repeated measurements were obtained from 4 slices for each region per mouse; Gnao1+/+: n=8, Gnao1+/G184S: n=7). 280! REFERENCES 281! REFERENCES Aglah, C., Gordon, T., & Posse de Chaves, E. I. (2008). cAMP promotes neurite outgrowth and extension through protein kinase A but independently of Erk activation in cultured rat motoneurons. Neuropharmacology, 55(1), 8-17. doi:10.1016/j.neuropharm.2008.04.005 Ananth, A. L., Robichaux-Viehoever, A., Kim, Y. M., Hanson-Kahn, A., Cox, R., Enns, G. M., . . . Bernstein, J. A. (2016). Clinical Course of Six Children With GNAO1 Mutations Causing a Severe and Distinctive Movement Disorder. Pediatr Neurol, 59, 81-84. doi:10.1016/j.pediatrneurol.2016.02.018 Andrade, R., Malenka, R. C., & Nicoll, R. A. (1986). A G protein couples serotonin and GABAB receptors to the same channels in hippocampus. Science, 234(4781), 1261-1265. Arnold, A. P., & Burgoyne, P. S. (2004). Are XX and XY brain cells intrinsically different? Trends Endocrinol Metab, 15(1), 6-11. Arya, R., Spaeth, C., Gilbert, D. L., Leach, J. L., & Holland, K. D. (2017). GNAO1-associated epileptic encephalopathy and movement disorders: c.607G>A variant represents a probable mutation hotspot with a distinct phenotype. Epileptic Disord, 19(1), 67-75. doi:10.1684/epd.2017.0888 Barhoumi, R., Qian, Y., Burghardt, R. C., & Tiffany-Castiglioni, E. (2010). Image analysis of Ca2+ signals as a basis for neurotoxicity assays: promises and challenges. Neurotoxicol Teratol, 32(1), 16-24. doi:10.1016/j.ntt.2009.06.002 Barkovich, A. J., Guerrini, R., Kuzniecky, R. I., Jackson, G. D., & Dobyns, W. B. (2012). for malformations of cortical 135(Pt 1348-1369. A developmental and genetic classification development: update Brain, doi:10.1093/brain/aws019 2012. 5), Beacham, D. W., Blackmer, T., M, O. G., & Hanson, G. T. (2010). Cell-based potassium ion channel screening using the FluxOR assay. J Biomol Screen, 15(4), 441-446. doi:10.1177/1087057109359807 282! Beckstead, M. J., & Williams, J. T. (2007). Long-term depression of a dopamine IPSC. J Neurosci, 27(8), 2074-2080. doi:10.1523/JNEUROSCI.3251-06.2007 Beyer, C., Kolbinger, W., Froehlich, U., Pilgrim, C., & Reisert, I. (1992). Sex differences of hypothalamic prolactin cells develop independently of the presence of sex steroids. Brain Res, 593(2), 253-256. doi:10.1016/0006-8993(92)91315-6 Beyer, C., Pilgrim, C., & Reisert, I. (1991). Dopamine content and metabolism in mesencephalic and diencephalic cell cultures: sex differences and effects of sex steroids. J Neurosci, 11(5), 1325-1333. Bleakman, D., Bowman, D., Bath, C. P., Brust, P. F., Johnson, E. C., Deal, C. R., . . . et al. (1995). Characteristics of a human N-type calcium channel expressed in HEK293 cells. Neuropharmacology, 34(7), 753-765. Brann, D. W., Dhandapani, K., Wakade, C., Mahesh, V. B., & Khan, M. M. (2007). Neurotrophic and neuroprotective actions of estrogen: basic mechanisms and clinical implications. Steroids, 72(5), 381-405. doi:10.1016/j.steroids.2007.02.003 Burgoyne, P. S. (1993). A Y-chromosomal effect on blastocyst cell number in mice. Development, 117(1), 341-345. Burgoyne, P. S., Thornhill, A. R., Boudrean, S. K., Darling, S. M., Bishop, C. E., & Evans, E. P. (1995). The genetic basis of XX-XY differences present before gonadal sex differentiation in the mouse. Philos Trans R Soc Lond B Biol Sci, 350(1333), 253-260 discussion 260-251. doi:10.1098/rstb.1995.0159 Cepeda, C., Murphy, K. P., Parent, M., & Levine, M. S. (2014). The role of dopamine in 235-254. Brain Res, 211, Huntington's doi:10.1016/B978-0-444-63425-2.00010-6 disease. Prog Chavez, C., Hollaus, M., Scarr, E., Pavey, G., Gogos, A., & van den Buuse, M. (2010). The effect of estrogen on dopamine and serotonin receptor and transporter levels in the brain: an autoradiography study. Brain Res, 1321, 51-59. doi:10.1016/j.brainres.2009.12.093 Chen, L. T., Gilman, A. G., & Kozasa, T. (1999). A candidate target for G protein action in brain. J Biol Chem, 274(38), 26931-26938. doi:10.1074/jbc.274.38.26931 283! Clepet, C., Schafer, A. J., Sinclair, A. H., Palmer, M. S., Lovell-Badge, R., & Goodfellow, P. N. (1993). The human SRY transcript. Hum Mol Genet, 2(12), 2007-2012. doi:10.1093/hmg/2.12.2007 Copenhaver, P. F., & Taghert, P. H. (1989). Development of the enteric nervous system in the moth. I. Diversity of cell types and the embryonic expression of FMRFamide-related neuropeptides. Dev Biol, 131(1), 70-84. Corey, S., & Clapham, D. E. (2001). The Stoichiometry of Gbeta gamma binding to G-protein-regulated inwardly rectifying K+ channels (GIRKs). J Biol Chem, 276(14), 11409-11413. doi:10.1074/jbc.M100058200 Czech, D. P., Lee, J., Sim, H., Parish, C. L., Vilain, E., & Harley, V. R. (2012). The human testis-determining factor SRY localizes in midbrain dopamine neurons and regulates multiple components of catecholamine synthesis and metabolism. J Neurochem, 122(2), 260-271. doi:10.1111/j.1471-4159.2012.07782.x Dewing, P., Chiang, C. W., Sinchak, K., Sim, H., Fernagut, P. O., Kelly, S., . . . Vilain, E. (2006). Direct regulation of adult brain function by the male-specific factor SRY. Curr Biol, 16(4), 415-420. doi:10.1016/j.cub.2006.01.017 Dhamija, R. M., J, W.; Shah, B, B.; Goodkin, H, P. (2016). GNAO1-Associated Movement Disorder. Movement Disorders Clinical Practice, 3(6), 615-617. doi:10.1002/mdc3.12344 Dietel, T. H., T. B.; Schlüter, G.; Cordes, I.; Trollmann, R.; Bast, T. (2016). Genotype and phenotype in GNAO1-Mutation – Case report of an unusual course of a childhood epilepsy. Neuropediatrics, 47(S 01), 1-17. doi:10.1055/s-0036-1583625 Dravid, S. M., & Murray, T. F. (2004). Spontaneous synchronized calcium oscillations in neocortical neurons in the presence of physiological [Mg(2+)]: involvement of AMPA/kainate and metabotropic glutamate receptors. Brain Res, 1006(1), 8-17. doi:10.1016/j.brainres.2004.01.059 Dunlap, K., & Fischbach, G. D. (1978). Neurotransmitters decrease the calcium ocmponent of sensory neurone action potentials. Nature, 276(5690), 837-839. Dupont, G., Combettes, L., Bird, G. S., & Putney, J. W. (2011). Calcium oscillations. Cold Spring Harb Perspect Biol, 3(3). doi:10.1101/cshperspect.a004226 284! Fabiato, A. (1983). Calcium-induced release of calcium from the cardiac sarcoplasmic reticulum. Am J Physiol, 245(1), C1-14. doi:10.1152/ajpcell.1983.245.1.C1 Farrell, M. R., Gruene, T. M., & Shansky, R. M. (2015). The influence of stress and gonadal hormones on neuronal structure and function. Horm Behav, 76, 118-124. doi:10.1016/j.yhbeh.2015.03.003 Feng, H., Khalil, S., Neubig, R. R., & Sidiropoulos, C. (2018). A mechanistic review on Dis. movement Neurobiol disorder. GNAO1-associated doi:10.1016/j.nbd.2018.05.005 Feng, H., Larrivee, C. L., Demireva, E. Y., Xie, H., Leipprandt, J. R., & Neubig, R. R. (2019). Mouse models of GNAO1-associated movement disorder: Allele- and sex-specific differences in phenotypes. PLoS One, 14(1), e0211066. doi:10.1371/journal.pone.0211066 Feng, H., Sjogren, B., Karaj, B., Shaw, V., Gezer, A., & Neubig, R. R. (2017). Movement disorder in GNAO1 encephalopathy associated with gain-of-function mutations. Neurology, 89(8), 762-770. doi:10.1212/WNL.0000000000004262 Frye, C. A. (2006). Role of androgens in epilepsy. Expert Rev Neurother, 6(7), 1061-1075. doi:10.1586/14737175.6.7.1061 Galea, L. A., Spritzer, M. D., Barker, J. M., & Pawluski, J. L. (2006). Gonadal hormone modulation of hippocampal neurogenesis in the adult. Hippocampus, 16(3), 225-232. doi:10.1002/hipo.20154 Garrett, J. E., & Wellman, C. L. (2009). Chronic stress effects on dendritic morphology in medial prefrontal cortex: sex differences and estrogen dependence. Neuroscience, 162(1), 195-207. doi:10.1016/j.neuroscience.2009.04.057 Georganta, E. M., Tsoutsi, L., Gaitanou, M., & Georgoussi, Z. (2013). delta-opioid receptor activation leads to neurite outgrowth and neuronal differentiation via a STAT5B-Galphai/o 329-341. doi:10.1111/jnc.12386 Neurochem, pathway. 127(3), J Goslin, K., Schreyer, D. J., Skene, J. H., & Banker, G. (1988). Development of neuronal polarity: GAP-43 distinguishes axonal from dendritic growth cones. Nature, 336(6200), 672-674. doi:10.1038/336672a0 285! Gould, E., Woolley, C. S., Frankfurt, M., & McEwen, B. S. (1990). Gonadal steroids regulate dendritic spine density in hippocampal pyramidal cells in adulthood. J Neurosci, 10(4), 1286-1291. Green, P. S., & Simpkins, J. W. (2000). Neuroprotective effects of estrogens: potential mechanisms of action. Int J Dev Neurosci, 18(4-5), 347-358. Greenough, W. T., Carter, C. S., Steerman, C., & DeVoogd, T. J. (1977). Sex differences in dentritic patterns in hamster preoptic area. Brain Res, 126(1), 63-72. doi:10.1016/0006-8993(77)90215-3 Harden, C., & MacLusky, N. J. (2004). Aromatase inhibition, testosterone, and seizures. Epilepsy Behav, 5(2), 260-263. doi:10.1016/j.yebeh.2003.12.001 Harden, C., & MacLusky, N. J. (2005). Aromatase inhibitors as add-on treatment for men 123-127. Neurother, 5(1), with Expert doi:10.1586/14737175.5.1.123 epilepsy. Rev Harte-Hargrove, L. C., Varga-Wesson, A., Duffy, A. M., Milner, T. A., & Scharfman, H. E. (2015). Opioid receptor-dependent sex differences in synaptic plasticity in the hippocampal mossy fiber pathway of the adult rat. J Neurosci, 35(4), 1723-1738. doi:10.1523/JNEUROSCI.0820-14.2015 He, J. C., Gomes, I., Nguyen, T., Jayaram, G., Ram, P. T., Devi, L. A., & Iyengar, R. receptor-mediated neurite (2005). The G alpha(o/i)-coupled cannabinoid outgrowth involves Rap regulation of Src and Stat3. J Biol Chem, 280(39), 33426-33434. doi:10.1074/jbc.M502812200 Herzog, A. G. (1991). Reproductive endocrine considerations and hormonal therapy for men with epilepsy. Epilepsia, 32 Suppl 6, S34-37. Heyser, C. J. (2004). Assessment of developmental milestones in rodents. Curr Protoc Neurosci, Chapter 8, Unit 8 18. doi:10.1002/0471142301.ns0818s25 Holahan, M. R. (2017). A Shift the Growth-Associated Protein (GAP-43) in the Coordination of Axonal Structural and Functional Plasticity. Front Cell Neurosci, 11, 266. doi:10.3389/fncel.2017.00266 to Supporting Role from a Pivotal for Holz, G. G. t., Rane, S. G., & Dunlap, K. (1986). GTP-binding proteins mediate 286! transmitter inhibition of voltage-dependent calcium channels. Nature, 319(6055), 670-672. doi:10.1038/319670a0 Horgan, A. M., & Copenhaver, P. F. (1998). G protein-mediated inhibition of neuronal migration requires calcium influx. J Neurosci, 18(11), 4189-4200. Horgan, A. M., Lagrange, M. T., & Copenhaver, P. F. (1994). Developmental expression of G proteins in a migratory population of embryonic neurons. Development, 120(4), 729-742. Huang, C. L., Jan, Y. N., & Jan, L. Y. (1997). Binding of the G protein betagamma subunit to multiple regions of G protein-gated inward-rectifying K+ channels. FEBS Lett, 405(3), 291-298. Hwangpo, T. A., Jordan, J. D., Premsrirut, P. K., Jayamaran, G., Licht, J. D., Iyengar, R., & Neves, S. R. (2012). G Protein-regulated inducer of neurite outgrowth (GRIN) modulates Sprouty protein repression of mitogen-activated protein kinase (MAPK) activation by growth factor stimulation. J Biol Chem, 287(17), 13674-13685. doi:10.1074/jbc.M111.320705 Ichikawa, M., Muramoto, K., Kobayashi, K., Kawahara, M., & Kuroda, Y. (1993). Formation and maturation of synapses in primary cultures of rat cerebral cortical cells: an electron microscopic study. Neurosci Res, 16(2), 95-103. Igarashi, M., Strittmatter, S. M., Vartanian, T., & Fishman, M. C. (1993). Mediation by G proteins of signals that cause collapse of growth cones. Science, 259(5091), 77-79. Ingalhalikar, M., Smith, A., Parker, D., Satterthwaite, T. D., Elliott, M. A., Ruparel, K., . . . Verma, R. (2014). Sex differences in the structural connectome of the human brain. Proc Natl Acad Sci U S A, 111(2), 823-828. doi:10.1073/pnas.1316909110 Jiang, M., & Bajpayee, N. S. (2009). Molecular mechanisms of go signaling. Neurosignals, 17(1), 23-41. doi:10.1159/000186688 Jiang, M., Gold, M. S., Boulay, G., Spicher, K., Peyton, M., Brabet, P., . . . Birnbaumer, L. (1998). Multiple neurological abnormalities in mice deficient in the G protein Go. Proc Natl Acad Sci U S A, 95(6), 3269-3274. doi:10.1073/pnas.95.6.3269 287! Jordan, J. D., He, J. C., Eungdamrong, N. J., Gomes, I., Ali, W., Nguyen, T., . . . Iyengar, R. (2005). Cannabinoid receptor-induced neurite outgrowth is mediated by Rap1 activation through G(alpha)o/i-triggered proteasomal degradation of Rap1GAPII. J Biol Chem, 280(12), 11413-11421. doi:10.1074/jbc.M411521200 Kehrl, J. M., Sahaya, K., Dalton, H. M., Charbeneau, R. A., Kohut, K. T., Gilbert, K., . . . Neubig, R. R. (2014). Gain-of-function mutation in Gnao1: a murine model of epileptiform encephalopathy (EIEE17)? Mamm Genome, 25(5-6), 202-210. doi:10.1007/s00335-014-9509-z Kight, K. E., & McCarthy, M. M. (2014). Using sex differences in the developing brain to identify nodes of influence for seizure susceptibility and epileptogenesis. Neurobiol Dis, 72 Pt B, 136-143. doi:10.1016/j.nbd.2014.05.027 Kulkarni, N., Tang, S., Bhardwaj, R., Bernes, S., & Grebe, T. A. (2016). Progressive Movement Disorder in Brothers Carrying a GNAO1 Mutation Responsive to Deep Brain 211-214. doi:10.1177/0883073815587945 Stimulation. 31(2), J Child Neurol, Larrivee, C. L., Feng, H., Leipprandt, J. R., Demireva, E. Y., Xie, H., & Neubig, R. R. (2019). Mice with GNAO1 R209H Movement Disorder Variant Display Hyperlocomotion Alleviated by Risperidone. bioRxiv. Latchney, S. E., Masiulis, I., Zaccaria, K. J., Lagace, D. C., Powell, C. M., McCasland, J. S., & Eisch, A. J. (2014). Developmental and adult GAP-43 deficiency in mice dynamically alters hippocampal neurogenesis and mossy fiber volume. Dev Neurosci, 36(1), 44-63. doi:10.1159/000357840 Lei, Q., Jones, M. B., Talley, E. M., Schrier, A. D., McIntire, W. E., Garrison, J. C., & Bayliss, D. A. (2000). Activation and inhibition of G protein-coupled inwardly rectifying potassium (Kir3) channels by G protein beta gamma subunits. Proc Natl Acad Sci U S A, 97(17), 9771-9776. doi:10.1073/pnas.97.17.9771 Livneh, Y., & Mizrahi, A. (2011). Experience-dependent plasticity of mature adult-born neurons. Nat Neurosci, 15(1), 26-28. doi:10.1038/nn.2980 Logothetis, D. E., Kurachi, Y., Galper, J., Neer, E. J., & Clapham, D. E. (1987). The beta gamma subunits of GTP-binding proteins activate the muscarinic K+ channel in heart. Nature, 325(6102), 321-326. doi:10.1038/325321a0 288! Lotto, B., Upton, L., Price, D. J., & Gaspar, P. (1999). Serotonin receptor activation enhances neurite outgrowth of thalamic neurones in rodents. Neurosci Lett, 269(2), 87-90. Lubin, M. L., Reitz, T. L., Todd, M. J., Flores, C. M., Qin, N., & Xin, H. (2006). A nonadherent cell-based HTS assay for N-type calcium channel using calcium 3 dye. Assay Drug Dev Technol, 4(6), 689-694. doi:10.1089/adt.2006.4.689 Luscher, C., & Slesinger, P. A. (2010). Emerging roles for G protein-gated inwardly rectifying potassium (GIRK) channels in health and disease. Nat Rev Neurosci, 11(5), 301-315. doi:10.1038/nrn2834 Masuho, I., Mototani, Y., Sahara, Y., Asami, J., Nakamura, S., Kozasa, T., & Inoue, T. (2008). Dynamic expression patterns of G protein-regulated inducer of neurite outgrowth 1 (GRIN1) and its colocalization with Galphao implicate significant roles of Galphao-GRIN1 signaling in nervous system. Dev Dyn, 237(9), 2415-2429. doi:10.1002/dvdy.21686 Mayer, A., Lahr, G., Swaab, D. F., Pilgrim, C., & Reisert, I. (1998). The Y-chromosomal genes SRY and ZFY are transcribed in adult human brain. Neurogenetics, 1(4), 281-288. McCool, B. A., Pin, J. P., Brust, P. F., Harpold, M. M., & Lovinger, D. M. (1996). Functional coupling of rat group II metabotropic glutamate receptors to an omega-conotoxin GVIA-sensitive calcium channel in human embryonic kidney 293 cells. Mol Pharmacol, 50(4), 912-922. Menke, L. A., Engelen, M., Alders, M., Odekerken, V. J., Baas, F., & Cobben, J. M. (2016). Recurrent GNAO1 Mutations Associated With Developmental Delay and a Movement 1598-1601. doi:10.1177/0883073816666474 Disorder. 31(14), J Child Neurol, Mize, A. L., & Alper, R. H. (2000). Acute and long-term effects of 17beta-estradiol on G(i/o) coupled neurotransmitter receptor function in the female rat brain as assessed by agonist-stimulated [35S]GTPgammaS binding. Brain Res, 859(2), 326-333. doi:10.1016/s0006-8993(00)01998-3 Murphy, T. H., Blatter, L. A., Wier, W. G., & Baraban, J. M. (1992). Spontaneous synchronous synaptic calcium transients in cultured cortical neurons. J Neurosci, 12(12), 4834-4845. 289! Nakamura, K., Kodera, H., Akita, T., Shiina, M., Kato, M., Hoshino, H., . . . Saitsu, H. (2013). De Novo mutations in GNAO1, encoding a Galphao subunit of heterotrimeric G proteins, cause epileptic encephalopathy. Am J Hum Genet, 93(3), 496-505. doi:10.1016/j.ajhg.2013.07.014 Nakata, H., & Kozasa, T. (2005). Functional characterization of Galphao signaling through G protein-regulated inducer of neurite outgrowth 1. Mol Pharmacol, 67(3), 695-702. doi:10.1124/mol.104.003913 Nash, M. S., Schell, M. J., Atkinson, P. J., Johnston, N. R., Nahorski, S. R., & Challiss, R. A. (2002). Determinants of metabotropic glutamate receptor-5-mediated Ca2+ and inositol 1,4,5-trisphosphate oscillation frequency. Receptor density versus agonist 35947-35960. doi:10.1074/jbc.M205622200 concentration. 277(39), Chem, Biol J Newberry, N. R., & Nicoll, R. A. (1985). Comparison of the action of baclofen with gamma-aminobutyric acid on rat hippocampal pyramidal cells in vitro. J Physiol, 360, 161-185. doi:10.1113/jphysiol.1985.sp015610 Ngun, T. C., Ghahramani, N., Sanchez, F. J., Bocklandt, S., & Vilain, E. (2011). The genetics of sex differences in brain and behavior. Front Neuroendocrinol, 32(2), 227-246. doi:10.1016/j.yfrne.2010.10.001 Okae, H., & Iwakura, Y. (2010). Neural tube defects and impaired neural progenitor cell in Gbeta1-deficient mice. Dev Dyn, 239(4), 1089-1101. proliferation doi:10.1002/dvdy.22256 Pacico, N., & Mingorance-Le Meur, A. (2014). New in vitro phenotypic assay for epilepsy: fluorescent measurement of synchronized neuronal calcium oscillations. PLoS One, 9(1), e84755. doi:10.1371/journal.pone.0084755 Paik, S., Somvanshi, R. K., & Kumar, U. (2019). Somatostatin-Mediated Changes in Microtubule-Associated Proteins and Retinoic Acid-Induced Neurite Outgrowth in SH-SY5Y 120-134. doi:10.1007/s12031-019-01291-2 Neurosci, 68(1), Cells. Mol J Pang, T., Atefy, R., & Sheen, V. (2008). Malformations of cortical development. Neurologist, 14(3), 181-191. doi:10.1097/NRL.0b013e31816606b9 Parducz, A., Hajszan, T., Maclusky, N. J., Hoyk, Z., Csakvari, E., Kurunczi, A., . . . 290! Leranth, C. (2006). Synaptic remodeling induced by gonadal hormones: neuronal plasticity as a mediator of neuroendocrine and behavioral responses to steroids. Neuroscience, 138(3), 977-985. doi:10.1016/j.neuroscience.2005.07.008 Peleg, S., Varon, D., Ivanina, T., Dessauer, C. W., & Dascal, N. (2002). G(alpha)(i) controls the gating of the G protein-activated K(+) channel, GIRK. Neuron, 33(1), 87-99. Pelosi, A., Menardy, F., Popa, D., Girault, J. A., & Herve, D. (2017). Heterozygous Gnal Mice Are a Novel Animal Model with Which to Study Dystonia Pathophysiology. J Neurosci, 37(26), 6253-6267. doi:10.1523/JNEUROSCI.1529-16.2017 Phan, A., Gabor, C. S., Favaro, K. J., Kaschack, S., Armstrong, J. N., MacLusky, N. J., & Choleris, E. (2012). Low doses of 17beta-estradiol rapidly improve learning and increase hippocampal dendritic spines. Neuropsychopharmacology, 37(10), 2299-2309. doi:10.1038/npp.2012.82 Reddy, D. S., & Jian, K. (2010). The testosterone-derived neurosteroid androstanediol is a positive allosteric modulator of GABAA receptors. J Pharmacol Exp Ther, 334(3), 1031-1041. doi:10.1124/jpet.110.169854 Reinoso, B. S., Undie, A. S., & Levitt, P. (1996). Dopamine receptors mediate differential morphological effects on cerebral cortical neurons in vitro. J Neurosci Res, 43(4), 439-453. doi:10.1002/(SICI)1097-4547(19960215)43:4<439::AID-JNR5>3.0.CO;2-G Reuveny, E., Slesinger, P. A., Inglese, J., Morales, J. M., Iniguez-Lluhi, J. A., Lefkowitz, R. J., . . . Jan, L. Y. (1994). Activation of the cloned muscarinic potassium channel by G protein beta gamma subunits. Nature, 370(6485), 143-146. doi:10.1038/370143a0 Rimler, A., Jockers, R., Lupowitz, Z., & Zisapel, N. (2007). Gi and RGS proteins provide biochemical control of androgen receptor nuclear exclusion. J Mol Neurosci, 31(1), 1-12. Robinson, H. P., Kawahara, M., Jimbo, Y., Torimitsu, K., Kuroda, Y., & Kawana, A. (1993). Periodic synchronized bursting and intracellular calcium transients elicited by low magnesium in cultured cortical neurons. J Neurophysiol, 70(4), 1606-1616. doi:10.1152/jn.1993.70.4.1606 291! Robinson, H. P., Torimitsu, K., Jimbo, Y., Kuroda, Y., & Kawana, A. (1993). Periodic bursting of cultured cortical neurons in low magnesium: cellular and network mechanisms. Jpn J Physiol, 43 Suppl 1, S125-130. Saitsu, H., Fukai, R., Ben-Zeev, B., Sakai, Y., Mimaki, M., Okamoto, N., . . . Matsumoto, N. (2016). Phenotypic spectrum of GNAO1 variants: epileptic encephalopathy to involuntary movements with severe developmental delay. Eur J Hum Genet, 24(1), 129-134. doi:10.1038/ejhg.2015.92 Sanger, T. D., Chen, D., Fehlings, D. L., Hallett, M., Lang, A. E., Mink, J. W., . . . Valero-Cuevas, F. (2010). Definition and classification of hyperkinetic movements in childhood. Mov Disord, 25(11), 1538-1549. doi:10.1002/mds.23088 Schmidt, C. J., Zubiaur, M., Valenzuela, D., Neer, E. J., & Drager, U. C. (1994). G(O), a guanine nucleotide binding protein, is expressed during neurite extension in the embryonic mouse. J Neurosci Res, 38(2), 182-187. doi:10.1002/jnr.490380208 Schorling, D. C., Dietel, T., Evers, C., Hinderhofer, K., Korinthenberg, R., Ezzo, D., . . . Kirschner, J. (2017). Expanding Phenotype of De Novo Mutations in GNAO1: Four New Cases and Review of Literature. Neuropediatrics, 48(5), 371-377. doi:10.1055/s-0037-1603977 Scott, R. H., & Dolphin, A. C. (1986). Regulation of calcium currents by a GTP analogue: potentiation of (-)-baclofen-mediated inhibition. Neurosci Lett, 69(1), 59-64. Sierra-Fonseca, J. A., Najera, O., Martinez-Jurado, J., Walker, E. M., Varela-Ramirez, A., Khan, A. M., . . . Roychowdhury, S. (2014). Nerve growth factor induces neurite outgrowth of PC12 cells by promoting Gbetagamma-microtubule interaction. BMC Neurosci, 15, 132. doi:10.1186/s12868-014-0132-4 Skucas, V. A., Duffy, A. M., Harte-Hargrove, L. C., Magagna-Poveda, A., Radman, T., Chakraborty, G., . . . Scharfman, H. E. (2013). Testosterone depletion in adult male rats increases mossy fiber transmission, LTP, and sprouting in area CA3 of hippocampus. 2338-2355. doi:10.1523/JNEUROSCI.3857-12.2013 Neurosci, 33(6), J Stephens, G. J. (2009). G-protein-coupled-receptor-mediated presynaptic inhibition in 421-430. Pharmacol Trends 30(8), cerebellum. the doi:10.1016/j.tips.2009.05.008 Sci, 292! Streb, H., Irvine, R. F., Berridge, M. J., & Schulz, I. (1983). Release of Ca2+ from a by nonmitochondrial inositol-1,4,5-trisphosphate. Nature, 306(5938), 67-69. intracellular store in pancreatic acinar cells Strittmatter, S. M., Fankhauser, C., Huang, P. L., Mashimo, H., & Fishman, M. C. (1995). Neuronal pathfinding is abnormal in mice lacking the neuronal growth cone protein GAP-43. Cell, 80(3), 445-452. Strittmatter, S. M., Fishman, M. C., & Zhu, X. P. (1994). Activated mutants of the alpha subunit of G(o) promote an increased number of neurites per cell. J Neurosci, 14(4), 2327-2338. Strittmatter, S. M., Igarashi, M., & Fishman, M. C. (1994). GAP-43 amino terminal peptides modulate growth cone morphology and neurite outgrowth. J Neurosci, 14(9), 5503-5513. Strittmatter, S. M., Valenzuela, D., Kennedy, T. E., Neer, E. J., & Fishman, M. C. (1990). G0 is a major growth cone protein subject to regulation by GAP-43. Nature, 344(6269), 836-841. doi:10.1038/344836a0 Strittmatter, S. M., Valenzuela, D., Sudo, Y., Linder, M. E., & Fishman, M. C. (1991). An intracellular guanine nucleotide release protein for G0. GAP-43 stimulates isolated alpha subunits by a novel mechanism. J Biol Chem, 266(33), 22465-22471. Strittmatter, S. M., Valenzuela, D., Vartanian, T., Sudo, Y., Zuber, M. X., & Fishman, M. C. (1991). Growth cone transduction: Go and GAP-43. J Cell Sci Suppl, 15, 27-33. Tanaka, M., Treloar, H., Kalb, R. G., Greer, C. A., & Strittmatter, S. M. (1999). G(o) protein-dependent survival of primary accessory olfactory neurons. Proc Natl Acad Sci U S A, 96(24), 14106-14111. doi:10.1073/pnas.96.24.14106 Thomas, P., Pang, Y., Dong, J., Groenen, P., Kelder, J., de Vlieg, J., . . . Tubbs, C. (2007). Steroid and G protein binding characteristics of the seatrout and human progestin membrane receptor alpha subtypes and their evolutionary origins. Endocrinology, 148(2), 705-718. doi:10.1210/en.2006-0974 Thompson, T. L., & Certain, M. E. (2005). Estrogen mediated inhibition of dopamine transport in the striatum: regulation by G alpha i/o. Eur J Pharmacol, 511(2-3), 293! 121-126. doi:10.1016/j.ejphar.2005.02.005 Thornhill, A. R., & Burgoyne, P. S. (1993). A paternally imprinted X chromosome retards the development of the early mouse embryo. Development, 118(1), 171-174. Traina, G., Petrucci, C., Gargini, C., & Bagnoli, P. (1998). Somatostatin enhances neurite outgrowth in PC12 cells. Brain Res Dev Brain Res, 111(2), 223-230. Tsien, R. W., Bean, B. P., Hess, P., Lansman, J. B., Nilius, B., & Nowycky, M. C. (1986). Mechanisms of calcium channel modulation by beta-adrenergic agents and dihydropyridine calcium agonists. J Mol Cell Cardiol, 18(7), 691-710. Vajda, F. J., Donnan, G. A., Phillips, J., & Bladin, P. F. (1981). Human brain, plasma, and cerebrospinal fluid concentration of sodium valproate after 72 hours of therapy. Neurology, 31(4), 486-487. doi:10.1212/wnl.31.4.486 VanDongen, A. M., Codina, J., Olate, J., Mattera, R., Joho, R., Birnbaumer, L., & Brown, A. M. (1988). Newly identified brain potassium channels gated by the guanine nucleotide binding protein Go. Science, 242(4884), 1433-1437. Vivar, C., Peterson, B. D., & van Praag, H. (2016). Running rewires the neuronal network 29-41. cells. Neuroimage, 131, adult-born granule of doi:10.1016/j.neuroimage.2015.11.031 dentate Wang, X., & Gruenstein, E. I. (1997). Mechanism of synchronized Ca2+ oscillations in 239-249. 767(2), Res, cortical Brain doi:10.1016/s0006-8993(97)00585-4 neurons. Wonders, C. P., & Anderson, S. A. (2006). The origin and specification of cortical interneurons. Nat Rev Neurosci, 7(9), 687-696. doi:10.1038/nrn1954 Worley, P. F., Baraban, J. M., Van Dop, C., Neer, E. J., & Snyder, S. H. (1986). Go, a guanine nucleotide-binding protein: immunohistochemical localization in rat brain resembles distribution of second messenger systems. Proc Natl Acad Sci U S A, 83(12), 4561-4565. doi:10.1073/pnas.83.12.4561 Xiong, J., Peng, J., Duan, H. L., Chen, C., Wang, X. L., Chen, S. M., & Yin, F. (2018). [Recurrent convulsion and pulmonary infection complicated by psychomotor retardation in an infant]. Zhongguo Dang Dai Er Ke Za Zhi, 20(2), 154-157. 294! Yang, H., Wan, L., Song, F., Wang, M., & Huang, Y. (2009). Palmitoylation modification of Galpha(o) depresses its susceptibility to GAP-43 activation. Int J Biochem Cell Biol, 41(7), 1495-1501. doi:10.1016/j.biocel.2008.12.011 Yeyeodu, S. T., Witherspoon, S. M., Gilyazova, N., & Ibeanu, G. C. (2010). A rapid, inexpensive high throughput screen method for neurite outgrowth. Curr Chem Genomics, 4, 74-83. doi:10.2174/1875397301004010074 Zamponi, G. W., Striessnig, J., Koschak, A., & Dolphin, A. C. (2015). The Physiology, Pathology, and Pharmacology of Voltage-Gated Calcium Channels and Their Future 821-870. doi:10.1124/pr.114.009654 Pharmacol Rev, 67(4), Therapeutic Potential. Zhang, S. P., Kauffman, J., Yagel, S. K., & Codd, E. E. (2006). High-throughput screening for N-type calcium channel blockers using a scintillation proximity assay. J Biomol Screen, 11(6), 672-677. doi:10.1177/1087057106289210 Zurawski, Z., Rodriguez, S., Hyde, K., Alford, S., & Hamm, H. E. (2016). Gbetagamma Binds to the Extreme C Terminus of SNAP25 to Mediate the Action of Gi/o-Coupled G Protein-Coupled Receptors. Mol Pharmacol, 89(1), 75-83. doi:10.1124/mol.115.101600 295!