FRAGILE X SYNDROME AND DIABETIC RETINOPATHY: PATHOLOGICAL CONDITIONS AND ASSOCIATE ALTERATION IN NEURONAL PROPERTIES By Hayyaf Saad Aldossary A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of Physiology—Master of Science 2017 ABSTRACT FRAGILE X SYNDROME AND DIABETIC RETINOPATHY: PATHOLOGICAL CONDITIONS AND ASSOCIATE ALTERATION IN NEURONAL PROPERTIES By Hayyaf Saad Aldossary Fragile X syndrome (FXS) and diabetic retinopathy are poorly treated conditions that dramatically affect patient’s and family members’ life style. FXS is the most common known form of inherited mental retardation. FXS is caused by a genetic mutation leading to decreased fragile X mental retardation protein (FMRP) production. The absence of FMRP leads to alterations in synaptic plasticity, which are dependent on activation of metabotropic glutamate receptor (mGluR) activation. In this study, short term activation of group I and II mGluRs is not altered between an animal model of FXS, Fmr1 knock out mice and wild-type mice. These negative findings suggest that short-lasting actions of mGluR activation in the neocortex may not contribute to the cognitive or sensory processing alterations associated with FXS. Diabetic retinopathy is a common complication of diabetes and is the leading cause for blindness in US working age adults. In this experiment, intraocular injection of the proinflammatory cytokine, interleukin 1-β, was used to mimic inflammation similar to that which occurs during diabetic retinopathy. We used electrophysiological recording techniques to determine the impact of this manipulation on the excitability of thalamocortical neurons in the dorsal lateral geniculate nucleus. We found alterations in excitability, which could lead to altered visual processing as identified in diabetic retinopathy. To my family iii ACKNOWLEDGMENTS I would like to send my gratitude for Dr. Cox, who directed me and supervised the research in this thesis, for his support and patience since my first day in his lab, I would also want to send my gratitude to Dr. Beatty, who had been a great mentor, and who taught me many things in the lab, without their help, support and positiveness, it would be impossible for me to get to where I am. I also want to take the chance to send my gratitude to all my lab mates; Dawn, Jacky, Kathleen, Laura, Bronson and all undergraduate students in the lab, who shared the moments of happiness, joy, and sometimes sadness with me. They have been a wonderful support to me; they helped me and directed me. Thank you all and I would not be able to succeed without your help. After this outstanding experience in my life, beside the knowledge I gained and the experience I earned, I am so grateful to become a member of a great family, the Cox lab family. iv TABLE OF CONTENTS LIST OF TABLES……………………………..………………………………………………………………….………………………vii LIST OF FIGURES………………………………………….…………………………………………………….…………………….viii KEY TO ABBREVIATIONS…………………………………….………………………………………………………….……….….x Chapter 1…..…………………………………………………………………………………………………..………………….………1 Literature Review………………………………………………..……………………….……………………….…….…1 Fragile X Syndrome ………………………………………………….………………………………………….……..…1 Introduction...……………………………………………………………….……..……………………………….1 FMR1 gene in Fragile X Syndrome..……….….…………………………….……….……………….…..2 The Significance of FMRP loss in Fragile X Syndrome...……….……….………..……....….….2 Behavior Phenotype in Fragile X Syndrome……..…..……………….………………………….…..3 Pathophysiology in Fragile X Syndrome...…………….…….……….………….……………………..4 Altered neuronal function associated with Fragile X Syndrome..….….…..……………….5 Hippocampus (mGluR theory)…………………………………….………….…………….…5 Neocortex………………………………………………………………….………….………………..6 Altered intrinsic properties.………………………..………….…...……………..6 Altered synaptic plasticity …………………………………….…...……………….7 Metabotropic Glutamate Receptors.…………………………………………………….….….….……….……8 Introduction ……………………………………………………………………………………………..…….….…8 mGluRs subtypes and Signaling pathways ….….………..……………….……...…………………..8 mGluRs and their role in Fragile X Syndrome.……...……………….….….……….……………..10 Diabetic Retinopathy………………………………………………………………………...……..…….…………...11 Introduction……………………………………………………………………………….………..……………..11 Stages of Diabetic Retinopathy……………………………………………….…………...……………..12 Non-Proliferative Diabetic Retinopathy………………………….…….……………….12 Proliferative Diabetic Retinopathy..………………….……………....….……………...12 The role of inflammation in Diabetic Retinopathy..……………………..….…………………..13 Interleukin 1 Beta (IL-1β) Pathway..………….….………………………………………..14 IL-1β and Diabetic Retinopathy.……………………………………………..……………..16 Alteration in neuronal function in diabetic retinopathy.………………….………………….17 The Thalamus……………………..…….………………………………………………….….…..…..…..…………..18 Introduction..………………………………………………………………………..………………..…....……18 The Lateral Geniculate Nucleus (LGN) and vision….….…..…………….….…..…..…………19 LGN Structure and cell types………………..…………………………………….……..…..…………..19 REFERENCES……………………………………………………………………….….……...…..………………21 v Chapter 2…………………………………………………………………………………………………………………………………31 Postsynaptic actions of mGluR activation in neocortical neurons are unaltered in Fmr1 KO mice….….……………..……………….….………………………………………….………………….…………….31 Abstract….…………………………………….……………………………………………………………….……………31 Introduction…….…………………………….……………………………………….………………….…..…..……..32 Materials and Methods…….…………….………………………………………………….……….….………….35 Results……….…………………………………….…………………………………………….……………..……………37 Group I mGluRs ………..……………………………………………………………….………………….…38 Group II mGluRs …………..……………………………………………………………….…………………40 Discussion……………………………………….………………………………………….…..…………………..…….42 REFERENCES.………………………………….…………………………………………….……….…………………..44 Chapter 3………………………………………………………………………………………….…………………………………..…48 Intraocular injection of Interleukin 1β alters intrinsic properties of dorsal lateral geniculate nucleus relay neurons.………..…………………………………….………………………...….48 Abstract….……………………………………………………………………………………………………………..….48 Introduction……………………………………………………………………………………..…….………….....…49 Materials and Methods…………………………………………………………….………..………………..…..52 Results……………………………………………………………………………………………..……………………….55 Discussion……………………………………………………………………………………..…………………..……..67 REFERENCES.……………………..………………………………….……………………….…………………………69 Chapter 4…………………………………………………………………………………………………………………………………72 Summary and future direction…..….………..………………………………….….…….……………..…..72 Summary.………….………………………………………………………………………………….….………………72 Future directions after the postsynaptic actions of mGluR in Fmr1 KO mice study..74 Future direction after the IL-1β intraocular injection and its alteration on the dorsal lateral geniculate nucleus relay neurons..…………….………..…….……………….…..…..75 vi LIST OF TABLES Table 2.1. Intrinsic properties of layer V/VI pyramidal neurons from WT and Fmr1 KO mice…............................................................................................................................................37 vii LIST OF FIGURES Figure 2.1. Group I mGluR agonist DHPG produces similar membrane depolarizations in pyramidal neurons from WT and Fmr1 KO animals. A. Voltage recordings from representative WT and Fmr1 KO neurons. DHPG produced a concentration dependent increase in amplitude in both types of neurons. B. Summary of population data (mean ± SEM).…….…….........................39 Figure 2.2. Group II mGluR agonist APDC produces similar membrane hyperpolarizations in pyramidal neurons from WT and Fmr1 KO animals. A. Voltage recordings from representative WT and Fmr1 KO neurons. APDC produced a concentration dependent increase in amplitude. B. Summary of population data (mean ± SEM)……………………………………………………………….….………..41 Figure 3.1. The relationship between intraocular IL-1β injection and resting membrane potential of dLGN relay neurons. (Top) Population graph of the average resting membrane potentials measured at various time points following IL-1β injection. 72 hours post-injection had a significantly hyperpolarized average resting membrane potential when compared to control mice (Cont.). No other time points tested were significantly different from control values. (Bottom) A representative trace showing where RMP was measured. *p<0.05…………..…….……56 Figure 3.2. The relationship between intraocular IL-1β injection and input resistance of dLGN relay neurons. (Top) Population graph of the average input resistance measured at various time points following IL-1β injection. No significant different was shown between time points tested and control values. (Bottom) A representative trace showing where Rin was measured………..58 Figure 3.3. The relationship between intraocular IL-1β injection and membrane time constant of dLGN relay neurons. (Top) Population graph of the average time constant measured at various time points following IL-1β injection. No significant alteration was shown between time points tested and control values. (Bottom) A representative trace showing how TC was measured…..60 Figure 3.4. The relationship between intraocular IL-1β injection and the average action potential discharge rate of dLGN relay neurons per second. (Top) Population graph of the average action potential discharge rate measured at various time points following IL-1β injection. No significant alterations were shown between time points tested and control values. (Bottom) A representative trace showing how the average action potential discharge rate was measured…………………………………………………………………………..…………………………………………………….62 Figure 3.5. The relationship between intraocular IL-1β injection and spike frequency adaptation (SFA) of dLGN relay neurons. (Top) Population graph of the average SFA measured at various time points following IL-1β injection. No significant alteration was shown between time points tested and control values. (Bottom) A representative trace showing how SFA was measured…64 Figure 3.6. The relationship between intraocular IL-1β injection and rheobase of dLGN relay neurons. (Top) Population graph of the average rheobase measured at various time points viii following IL-1β injection. No significant alteration was shown between time points tested and control values. (Bottom) A representative trace showing how rheobase was measured……….…66 ix KEY TO ABBREVIATIONS ADHD: Attention defect hypersensitivity disorder ASD: Autism spectrum disorder BM: basement membrane dLGN: Dorsal lateral geniculate nucleus FMR1: Fragile X mental retardation gene FMRP: Fragile X mental retardation protein FXS: Fragile X syndrome IGL: Intergeniculate nucleus IL: Interleukin IL-1β: Interleukin-1 beta ILR: Interleukin receptor KO: Knocked out LGN: Lateral geniculate nucleus LTD: Long-term depression LTP: Long-term potentiation mGluRs: Metabotropic glutamate receptors MGN: Medial geniculate nucleus mRNA: Messenger RNA RGCs: Retinal ganglion cells RMP: Resting membrane potential Rin: Input resistance SEM: Standard error of the mean x vLGN: Ventral lateral geniculate nucleus VPN: Ventral posterior nucleus WT: Wildtype NPDR: Non-proliferative diabetic retinopathy PDR: Proliferative diabetic retinopathy xi Chapter 1 Literature Review Fragile X Syndrome Introduction Fragile X syndrome (FXS) is one of the most common known causes of inherited mental retardation, and it is associated with many mental defects including cognitive, physical and biological alteration (Crawford et al., 2001). FXS is a genetic disorder in which there is a significant reduction of the fragile X mental retardation protein (FMRP), which results from alteration in the fragile x mental retardation (FMR1) gene (Morgan, 1911; Li et al., 2002; Ridaura-Ruiz et al., 2009). The decreased FMRP levels is responsible for the cognitive deficits, which can range from mild to severe alterations (Li et al., 2002; Ridaura-Ruiz et al., 2009). FXS affects approximately 1 in 4000 males and 1 in 8000 females and the cognitive/behavioral alterations includes multiple disorders including attention-deficit disorder, anxiety disorder, epilepsy, and delays in speech and language development (Sabaratnam, 2006). The lack of FMRP expression plays a major role in altering the normal development of synaptic connectivity, synaptic transmission, as well as alterations in the intrinsic properties of neurons (Pfeiffer and Huber, 2009; Zhang et al., 2016; Wilson and Cox, 2007). In addition to neurological deficits, many FXS patients also have altered physical characteristics including long and narrow face, large forehead, and large ears (Ridaura-Ruiz et al., 2009). 1 FMR1 gene in Fragile X Syndrome FXS is a disorder caused by genetic mutation. In FXS, the X chromosome specifically in the 5’ untranslated region, there is an increase in the CGG repeat times; normally range from (1 to 50 CGG repeats), the permutation range is 50-200 CGG repeats, and when the number of CGG repeats exceeds 200, as seen in FXS, the FMR1 gene will be silenced. This mutation leads to the loss of the FMRP production (Rousseau at el., 1994). The permutation in FMR1 gene (50200 repeats) results in abnormal FMRP production (Brouwer et al., 2009), and can lead to a couple of fragile X related conditions: fragile X associated primary ovarian insufficiency and fragile X-associated tremor syndrome. Fragile X associated primary ovarian insufficiency is a condition that can lead to infertility or in mild conditions cause a reduced fertility. On the other hand, fragile X-associated tremor syndrome is a condition that affects both men and women, and is characterized by tremors, memory loss, behavior changes and peripheral neuropathy and can be misinterpreted as Parkinson’s disease (Sheridan et al., 2011). The significance of FMRP loss in Fragile X Syndrome It is generally thought that the reduction or absence of FMRP is the cause for the broad range of deficits/alterations associated with FXS (Abrams et al., 1997). FMRP is an RNA binding protein that influences a large number of mRNAs including many that are found in the brain (Darnell and Klann, 2013; Bardoni et al., 1999). FMRP plays a significant role in local protein synthesis at the postsynaptic site, it is also crucial for dendritic spine maturation (Schenck et al., 2001). FMRP could also regulate and maintain long-lasting changes in synaptic strength, due to its role in local protein synthesis (Kim et al., 2009). The loss of FMRP production will lead to the 2 reduction of synaptic plasticity and the developmental spine maturation, contributing in the cognitive defects and dendritic spines immature associated with FXS (Sidorov et al., 2013; Irwin et al., 2002). FMRP is also crucial for the translation and transportation of the mRNA, which makes its loss contributes to the alteration in the translation signaling pathway (Bassell and Warren, 2008). It has been showed that this alteration is leading to defects in synaptic function, which in turn causes behavioral phenotype in individuals with FXS (Bassell and Warren, 2008). Behavioral phenotype in Fragile X Syndrome The behavioral phenotype associated with FXS involves cognitive disability, learning and language difficulties, and social anxiety (Huddleston et al., 2014). Most patients with FXS present with developmental delay or intellectual disability, and considering FXS is associated with an X chromosome, male patients tend to have higher severity in the manifestation of behavioral defects (Gallagher A and Hallahan, 2012). It has also been shown that males have a lower IQ range (40-70), while heterozygotes females with FXS have an IQ range of (70-90; Hagerman et al., 1992; Huddleston et al., 2014). Individuals with FXS typically show other behavioral disorders, including attention defect hypersensitivity disorder (ADHD), which is seen in approximately 75% of FXS patients. It also have been reported that FXS is associated with increased anxiety, which is displayed by behaviors such as isolation and shyness, is accounted for approximately 65% of FXS patients (Boyle and Kaufmann, 2010). Almost half of individuals with FXS were either diagnosed or treated for general hyperactivity, which is considered to be one of another common abnormal behaviors associated with FXS (Sullivan et al., 2006; Boyle and Kaufmann, 2010). Approximately 3 30% of individuals with FXS were also diagnosed with autism spectrum disorder (ASD; Boyle and Kaufmann, 2010). In addition, approximately 20% of FXS individuals are also diagnosed with a seizure disorder (Hagerman et al., 2009). These findings indicate that FXS is associated with many neurological defects that could play a significant role in the misleading of its diagnosis leading to late recognition of the condition. Pathophysiology in Fragile X Syndrome Thomas H. Morgan was first one to discover the sex-linked characteristics in 1911, when he saw the “white eye” phenomena affecting males only. He saw that in drosophila flies that have a red eye color originally, only males could have a white eyes color after breeding. Morgan then suggested that there could be genetic features that were sex-linked and thus would preferentially affect one sex (males) as opposed to females. Years after Morgan, studies showed that FXS occurrence is higher in males than females, 2:1 ratio (Turner et al., 1986; Pengarikano et al., 2007). Fmr1 KO mice are FXS mice models that were genetically modified by The Dutch and Belgian group, this modification led to the inactivation of Fmr1 gene causing the prevention of FMRP production (Bakker et al., 1994). After modification, the Fmr1 KO mice showed an increase in motor activity and in exploratory behaviors, which can also be seen in individuals with FXS (Bakker et al., 1994). This animal model provided an opportunity to study the underlying mechanisms that give rise to specific FXS phenotypes. By knowing more about the neurological alterations in Fmr1 KO mice, we may be able to develop appropriate treatments for FXS patients. 4 Altered neuronal function associated with Fragile X Syndrome In Fmr1 KO mice, the dendritic spines are abnormally immature due to the loss of FMRP (Martín et al., 2010). Anatomical studies have reported that dendritic spines in both FXS patients and Fmr1 KO mice look thinner and longer than normal (Martín et al., 2010). These changes in spine morphology are indicators for dendritic spine immaturity. These defects in dendritic spines lead to neuronal function abnormalities including synaptic plasticity alteration (Pfeiffer and Huber, 2009). Alterations in neuronal function associated with FXS have been studied in many brain regions. Here we will review alterations in hippocampal function that were found in Fmr1 KO mice and subsequently gave rise to the mGluR theory. Hippocampus (mGluR theory) In the hippocampus, it has been shown that the magnitude of long-term depression (LTD), which is the activity dependent reduction in the efficacy of the neuronal synapse, is enhanced in Fmr1 KO animals after the activation of group I mGluRs. The reason for this increase is that the absence of FMRP will lead to the exaggeration in group I mGluR dependent protein synthesis (Bear et al., 2004). The over-activation of the group I mGluR signaling might contributes to cognitive impairment, developmental delay and the loss of motor coordination in FXS (Bear et al., 2004).It has been reported by (Bear et al., 2004) that the increase in LTD in the hippocampus of FXS mice could slow the net synaptic maturation, by shifting the synaptic balance towards synaptic loss rather than synaptic gain, which will eventually lead to the cognitive defects and the development delay associated with FXS (Bear et al., 2004). 5 Neocortex Altered intrinsic properties The neocortex also exhibits alterations in neuronal function. In the somatosensory neocortex, FXS neurons show that the resting membrane potential (RMP) and the time constant (TC) did not change when comparing neonatal mice (postnatal age: 7 days) with older mice (postnatal age: 28 days), while when the same comparison was done on WT mice, both RMP and TC where altered. These alterations indicating that the loss of FMRP is slowing the neuronal maturation process, which could contribute in the defects associated with FXS (Zhang et al., 2016). It has also been reported that pyramidal cortical neurons from Fmr1 KO mice exhibit more intrinsic properties excitability than WT mice, which may contribute to seizure and seizure-like neuronal activity in FXS (Contractor et al., 2015). Another neuronal alteration is an altered action potential firing rate; the discharge rate of cortical pyramidal neurons is higher in Fmr1 KO neurons compared to WT neurons and the time for the first spike is shorter in Fmr1 KO neurons compared to WT mice, indicating hyperactivity that could lead to altered sensory processing (Contractor et al., 2015). And when the persistent activity states, or UP state, which occur predominantly during slow-wave sleep, was studied by Hays et al., they showed that the UP state is ~50% longer in the Fmr1 KO when compared with the WT mice. This finding supports that the Fmr1 KO mice exhibit more excitatory circuits than the WT mice (Hays et al., 2011). 6 Altered synaptic plasticity Synaptic plasticity is the ability of the neurons to change due to behavioral stimuli; i.e. strengthening (LTP) or weakening (LTD) changes (Berlucchi and Buchtel, 2009). In hippocampus, local protein synthesis in the presence of FMRP plays a significant role in synaptic plasticity, which is present in the dendritic spine, where synaptic transmission occurs (Pfeiffer and Huber, 2009). The absence of FMRP in Fmr1 KO mice contributes to the alteration in plasticity in Fmr1 KO neocortical neurons. It has been shown that in neocortex, Fmr1 KO mice exhibit significant alterations in LTP, due to the loss of FMRP (Wilson and Cox, 2007). It has also been shown that in deep layer neocortex, LTP is primarily dependent on mGluR5 activation (Wilson and Cox, 2007). Larson et al (2005) have suggested that in Fmr1 KO mice, LTP alteration could be age dependent, they showed that there was no alteration in LTP in Fmr1 KO mice younger than 6 months, after that the LTP showed significant alteration, especially 12 months post-natal age. 7 Metabotropic Glutamate Receptors Introduction Most excitatory neurons in the central nervous system use glutamate as their primary neurotransmitter (Tanabe et al., 1992). Glutamate receptors play significant roles in synaptic transmission, including synaptic plasticity, which is altered in Fmr1 KO mice (Monaghan et al., 1989). There are two families of glutamate receptors: ionotropic glutamate receptors, which are ligand-gated ion channels that mediate a fast synaptic transmission consisting of NMDA, AMPA and kainate receptors (Wisden and Seeburg, 1993). The other glutamate receptor family is the mGluRs, multiple subtypes of G-protein coupled receptors that mediate slow synaptic transmission and serve as neuromodulators (Ferraguti and Shigemoto, 2006). mGluRs subtypes and Signaling pathways There are multiple subtypes of mGluRs based on their function in the central nervous system (Conn and Pin, 1997). Based on their amino acid sequence, eight distinct mGluRs (mGluRs1-8) have been identified and these have been divided into three groups; group I mGluRs (including mGluR1 and mGluR5), group II mGluRs (including mGluR2 and mGluR3) and group III mGluRs (including mGluR4, mGluR6, mGluR7 and mGluR8), (Jeffrey and Pin, 1997). Group I mGluR: mGluR1 and mGluR5 are typically located postsynaptically. The activation of group I mGluRs will activate phospholipase C, resulting in the generation of IP3. This pathway will lead to the mobilization of calcium from the endoplasmic reticulum, which in turn cause a 8 depolarization response as well as the activation of protein kinase C (PKC; Niswender and Conn, 2010). Group II mGluR: mGluR2 and mGluR3, are located pre- and post-synaptically. The presynaptic activation of group II mGluRs will result in the inhibition of adenylyl cyclase and subsequent inhibition of cAMP as well as the inhibition of calcium channels (Bellonea et al., 2008; Niswender and Conn, 2010). On the other hand, the postsynaptic activation of group II mGluRs will result in the activation of the G-protein-coupled inward rectifying (GIRK) channels (Bellone et al., 2008). The activation of both pre and postsynaptic sites of group II mGluRs will lead to an inhibitory effect, which in turn will result in neuronal hyperpolarization (Niswender and Conn, 2010). One of their most selective agonist is 4-amino-2,4-pyrrolidinedicarboxylic acid (APDC; Niswender and Conn, 2010). Group III mGluR: mGluR4, mGluR6, mGluR7 and mGluR8, in group III mGluR, mGluR4, mGluR7 and mGluR8 are expressed in neurons, while mGluR6 is expressed in retinal bipolar cells (Nakajima et al., 1993). mGluR4, mGluR7 and mGluR8 are located presynaptically, while mGluR6 is located postsynaptically in retinal cells (Coutinho and Knöpfel, 2002) (Coutinho and Knöpfel, 2002). Signaling of mGluR4, mGluR7 and mGluR8 occurs through the inhibition of adenylyl cyclase, inhibition of cAMP, the activation of potassium channels, and the inhibition of calcium channels, while mGluR6 signaling occurs through the stimulation of cGMP phosphodiesterase, leading to neuronal hyperpolarization and inhibitory effect. (Niswender and Conn, 2010). 9 mGluRs and their role in Fragile X Syndrome mGluRs are involved in many impactful elements of brain function, such as neuronal plasticity, synaptic transmission and neuronal development (Coutinho and Knöpfel, 2002). Learning and memory is considered to be the most researched brain function and mGluRs play a significant role, while mGluR induced plasticity is crucial for cognitive functions (Riedel et al., 2003). One function of FMRP is to regulate mRNA translation and transportation; both are critical for synaptic maturation and function. (Pfeiffer and Huber, 2009). It has been shown by Antar et al. that in hippocampus, mGluRs activation increases FMRP localization in the dendritic spines, which is significant for synaptic plasticity, and the use of group I and II mGluR antagonist was sufficient to prevent this FMRP localization (Antar et al., 2004). Multiple subtypes of mGluRs play a significant role in the neocortex, where many brain functions including sensory perception and motor commands take place. Neuronal plasticity is being influenced by mGluRs; the activation of group I mGluRs contribute in LTP induction, which is attenuated in Fmr1 KO mice (Wilson and Cox, 2007). The absence of mGluR function in the neocortex would lead to neurological defects in plasticity after brain stimulation (Wilson and Cox, 2007). The involvement of LTP with the activation of mGluRs could contribute in the impairment in cognitive functions associated with FXS (Desai et al., 2006). 10 Diabetic Retinopathy Introduction Diabetic retinopathy is the leading cause of blindness in working-age adults and it is considered to be one of the more serious complications associated with diabetes mellitus (Kempen et al., 2004). The pathogenesis of diabetic retinopathy involves vascular, neuronal, and inflammatory processes (Qiu et al., 2016). In diabetic retinopathy, the inflammatory response will increase proinflammatory cytokine levels, including interleukin (IL)-1β, IL-6 and IL8, in the ocular fluid (Qiu et al., 2016). These inflammatory components will lead to retinal neovascularization, blood-retinal barrier breakdown, and eventually retinal ganglion apoptosis (Kempen et al., 2004; Qiu et al., 2016; Zhang et al., 2017). The early stage of the disease is the leading cause for visual defects and the late stage is the leading cause of blindness in working age adults (Zhang et al., 2017). Excitotoxicity of glutamate, the reduction of nerve growth factors, and inflammatory factor release are associated with diabetic retinopathy (Zhang et al., 2017). Retinal ganglion cell protection could be an effective way to prevent the pathogenesis of diabetic retinopathy (Zhang et al., 2017). In diabetic retinopathy, the retinal function is gradually reduced; it usually starts with a reduction in night vision and might continue worsening to a total vision loss (Caldwell et al., 2003). When diabetic retinopathy starts, alterations in retinal blood flow and RGC function will occur, as well as thickening of the basement membrane and an increase in vascular permeability (Caldwell et al., 2003). 11 Stages of diabetic retinopathy Non-Proliferative Diabetic Retinopathy Non-proliferative diabetic retinopathy is the early stage of diabetic retinopathy and is associated with vasodegeneration lesions in the retinal microvascular bed (Gardiner et al., 2007). This stage of diabetic retinopathy is characterized by the thickening of the capillary basement membrane (BM), vascular smooth muscle dropout, microanyorisms, and capillary occlusion (Gardiner et al., 2007). The first stage of diabetic retinopathy is associated with visual defects, such as dark spots vision and blindness is not a risk factor in the early stage, but if untreated will eventually lead to the advanced stage (Curtis et al., 2009). Since diabetic retinopathy mainly affects the retina, the most common symptom of diabetic retinopathy is retina capillary cell loss, which is caused by an increase in the number of microaneurysms over time. This loss in capillary cells will lead to the diabetic retinopathy visual associated defects (Cai and Boulton, 2002). The increase in vascular permeability, and the symptom associated with this stage will lead to the reduction in central accuracy caused by retinal trephining and edema (Antonetti et al., 2006). Proliferative Diabetic Retinopathy Proliferative diabetic retinopathy occurs when new vessels and connective tissues start to grow on the outer layer of the retina or the optic nerve (Kroll et al. 2007). This new growth is an indicator of retinal damage and it starts many years after the initiation of the nonproliferative stage (Kroll et al., 2007). In this advanced stage of diabetic retinopathy, retinal capillary aneurysms and the loss of perfusion of capillaries and arterioles take place (Engerman, 12 1989). There are some associated damage in the retina, such as vessel leakage and hemorrhage, which in turn will activate Muller cells and microglial cells and initiate an inflammatory process (Engerman, 1989). In this stage, neuronal impairments are the primary symptom, where the reduction in night and color vision, caused by nerve defects (Antonetti et al., 2006). Eventually, if not treated, diabetic retinopathy has a very high probability of causing blindness (Antonetti et al., 2006; Cai and Boulton, 2002). The role of inflammation in Diabetic Retinopathy Diabetic retinopathy is classified as an inflammatory disease, because of the increased number of leukocytes present at the retinal vasculature in diabetic patients after retinal vascular leakage, capillary non-perfusion, and endothelial cell damage (Joussen et al., 2004). The activation of Muller and microglia cells will trigger the inflammatory process; many pathways are activated including interleukin-1 beta (IL-1β), which induces the expression of proinflammatory proteins (Tang and Kern, 2011). The increase expression of proinflammatory proteins (IL-1β and others) will lead to the synthesis of more cytokines and chemokines (Tang and Kern, 2011). Most of the inflammatory changes such as the activation of Mullar cells and its downstream effect, are keys for diabetic retinopathy development, without these inflammatory changes, many characteristics of diabetic retinopathy will be blocked (Joussen et al., 2004). Interleukins are a large group of immunomodulatory proteins, which have many responses in cells and tissues (Akdis et al., 2011). They initiate immune responses by binding to high-affinity receptors on the cell surface (Mizel, 1989). Interleukins have many complex 13 functions including cell proliferation, maturation, migration, and adhesion (Akdis et al., 2011). The interleukins have many subtypes, where each one has its specific role and pathway and all of them work as a part of the immune system (Akdis et al., 2011; Mizel, 1989). The interleukin family consists of more than 35 subtypes and their structure is dimer, heterodimer, or monomer (Akdis et al., 2011). IL-1 and IL-17 are proinflammatory cytokines, while IL-23 stimulates the production of IL-17 (Akdis et al., 2011). IL-1 has two subtypes IL-1α and IL-1β and it has two receptors IL-1RI and IL-1RII and macrophages are their major source (Akdis et al., 2011). On the other hand, IL-17 has one receptor IL-17R and its main source is the T helper 17 cells (Akdis et al., 2011). Interleukin 1 Beta (IL-1β) Pathway The activation of IL-1 would contribute in many inflammatory conditions leading to several medical situations (Dinarello, 1996). IL-1β is activated by the caspase 1 gene, which is highly involved in cell proliferation, differentiation, and apoptosis processes and its activity could be influenced by gene expression modification (Dinarello, 1988). Activation of IL-1β will lead to activation of cyclooxygenase inducing inflammatory hypersensitivity (Dinarello, 1988; Weber et al., 2010). During systemic inflammation, an overexpression of IL-1β gene will occur in the central nervous system, due to its significance as a neuroregulator (Wong et al., 1997). In general, IL-1 plays a significant role in inflammatory process; IL-1α and IL-1β are synthesized as pro IL-1α and pro IL-1β, which can be cleaved to generate mature forms (Gabay 14 et al., 2010). Pro IL-1α can be found as a membrane-associated protein at many cell-type surfaces, which is involved in cell-to-cell signaling (Claudia, 2009). Pro IL-1β is inactive and it needs to be activated to function, its activation requires the activation of caspase 1, gram positive and gram negative bacteria could activate caspase 1 (Claudia, 2009). IL-1β release requires two signals from the primary microphages; the first signal induces transcription and translation, while the second signal activates caspase 1, these two signals will result in the production of active IL-1β (Gabay et al., 2010). Interleukin cytokines are associated with infection and inflammation, some cytokines promote inflammation, such as the proinflammatory cytokines, others suppress the activity of the proinflammatory type, which are the anti-inflammatory cytokines (Dinarello, 2000). Their function is based on gene coding and it is likely that the proinflammatory cytokines will not be produced in a healthy individual. These genes are targeted in conditions where immunity is already compromised, during infection or trauma for example (Dinarello, 2000). When the inflammatory cascade is triggered by the proinflammatory cytokines, it will induce nitric oxide synthesis, leading eventually to the activation of neutrophils and this is when inflammatory tissue destruction and loss of function occur. 15 IL-1β and Diabetic Retinopathy Many members of the interleukin family, including IL-1β, were detected in the vitreous fluid of patients with diabetic retinopathy and the expression of IL-1β is upregulated in retinal ischemic conditions (Kowluru and Odenbach, 2004). This indicates that the involvement of IL-1β cytokine may play a significant role in the pathogenesis of diabetic retinopathy. It was shown by Kowluru and Odenbach, 2004, that the increase of IL-1β expression in the retina showed 70% raise in cell death as well as the acceleration of retinal cell death processes. The expression of IL-1β is increased when retinal endothelial cells were exposed to high glucose, this increase in expression will lead to a dramatic increase in nitric oxide levels (Kowluru and Odenbach, 2004). The increase in IL-1β levels will significantly increase endothelial cells apoptosis (Kowluru and Odenbach, 2004; Abu El-Asrar, 2012). In conclusion, the continued hyperglycemia circulation is the main reason of keeping this positive feedback cycle going, leading to more activation of IL-1β, which in turn cause the development of retinopathy (Kowluru and Odenbach, 2004). 16 Alteration in neuronal function in diabetic retinopathy The advanced stage of diabetic retinopathy is associated with blindness; this blindness is mainly due to the increase in retinal ganglion cell (RGC) apoptosis (Vujosevic and Midena, 2013). It has been shown that diabetic retinopathy is not only a vasculature disease, it also involves neurons and glial cells in the retina (Vujosevic and Midena, 2013; Whitmire et al., 2011). After an acute retinal injury, glial cells will be activated to protect retinal neurons (Whitmire et al., 2011). The activation of glial cells will result in growth factors release, which will promote either cell survival or cell death (Whitmire et al., 2011). Retinal neurodegeneration occurs before the development of clinically detectable microvascular damage, indicating that retinal neurodegeneration plays a major role in the microvascular changes seen in diabetic retinopathy (Vujosevic and Midena, 2013). In diabetic retinopathy conditions, apoptosis will occur in different retinal layers and from different types of neurons, as well as the reduction in retinal thickness due to ganglion cells loss (Vujosevic and Midena, 2013; Whitmire et al., 2011). All of these alterations in the retina and RGCs could lead to alterations in its signaling recipient, which is the dLGN. 17 The Thalamus Introduction The thalamus is a midbrain structure that is often referred to as the major relay station to the neocortex. With regards to sensory systems, most peripheral sensory information passes through the thalamus in route to the neocortex. The thalamus is divided into three groups; specific, associated and non-specific nuclei (Herrero et al., 2002). The specific nuclei, which are connected to specific areas in the neocortex, are composed of three nuclei; the medial geniculate nucleus (MGN), the dorsal lateral geniculate nucleic (dLGN) and the ventral posterior nucleus (VPN) (Ward, 2013; Zhang, 1988). The MGN is the relay nucleus for the auditory system; it receives auditory input from brain stem auditory nuclei and sends signals to the auditory cortex (Ward, 2013). The dLGN is the visual sensory relay nucleus. It receives retinal output and sends visual information to the visual cortex (Ward, 2013), while the VPN is the somatosensory relay nucleus; it receives its input from cerebellum and projects to layer IV somatosensory cortex (Zhang, 1988). The non-specific nuclei, projects to non-specific areas and receives inputs from both the cortex and the thalamus, (Pinault, 2004; Swenson, 2006). It mainly projects to thalamic nuclei; its function plays a significant role in the regulation of thalamic input, which in turn will regulate thalamic output. The non-specific nuclei participate in the regulation of many high order functions in the brain (Pinault, 2004; Swenson, 2006). 18 The lateral geniculate nucleus and vision The lateral geniculate nucleus is divided functionally into three subdivisions; dorsal (dLGN), ventral (vLGN) and intergeniculate (IGL) (Harrington, 1997). The role of dLGN is to relay visual information from the retinal ganglion cells (RGCs) to primary visual cortex (Essen and Anderson, 1995; Dhande and Huberman, 2014). The function of the vLGN is not precisely known, but it has been suggested that it has roles in both brightness discrimination as well as the regulation of circadian rhythms (Harrington, 1997). The third division of the LGN is the IGL, which plays a significant role in the modification of circadian rhythms, it is also suggested that IGL contributes in the regulation of sleep and arousal and visuomotor functions (Harrington, 1997). In visual processing, and when visual information gets to the RGCs in the retina, it will be projected to the relay neurons of the dLGN. The dLGN in turn relays the visual information to the visual cortex (van Essen and Anderson, 1995; Huberman and Niell, 2011). Historically, the dLGN was thought to serve as a passive relay station, but over the last several decades it appears that there is significant processing and modification of visual information prior to going to visual cortex. LGN Structure and cell types The dLGN is the gateway between the retina and the visual cortex. In the dLGN, less than 10% of synapses come from the retina as a driving input, while the rest is an inhibitory input from visual cortex as well as the brain stem (Sherman and Guillery, 2002). The inhibitory influence in the dLGN controls many retinogeniculate transmissions, which in turn leads to the prevention 19 of hyperexcitation conditions manifested by burst and tonic mode in the relay neurons, which are two firing modes in the thalamic relay neurons (Sherman and Guillery, 2002). These two firing modes depend on the activation of T type calcium channels, when the T channels are inactivated due to depolarization; the tonic mode will occur (Sherman, 2001). On the other hand, hyperpolarization will activate T channels leading to the burst firing mode (Sherman, 2001). The dLGN relay neurons are divided into three groups; X (biconical), Y (symmetrical) and W (hemispherical), based on their neuronal properties and function (Sherman, 1985; Krahe et al., 2011; Tang et al., 2016). dLGN cell types function were described by Sherman in 1985 as: the Y cell type is responsible for the analysis of basic visual information, the X cell type will provide high resolution for the analyzed input, while the W cell type plays a role in conscious perception of visual patterns (Sherman, 1985). In the dLGN there are relay neurons and interneurons; the relay neurons are responsible for the thalamocortical communication (Dhande and Huberman, 2014), while the interneurons, which account for ~30% of dLGN neurons, function is to inhibit the thalamocortical neurons and modulate the signal transmission in the dLGN (Leist et al., 2016). 20 REFERENCES 21 REFERENCES Abrams MT, Doheny KF, Mazzocco MM, Knight SJ, Baumgardner TL, Freund LS, Davies KE and Reiss AL. Cognitive, behavioral, and neuroanatomical assessment of two unrelated male children expressing FRAXE. American Journal of Medical Genetics. 1997: 74, 73-81. Abu El-Asrar AM. Role of inflammation in the pathogenesis of diabetic retinopathy. Middle East African Journal of Ophthalmology. 2012: 19, 70-74. Akdis M, Burgler S, Crameri R, Eiwegger T, Fujita H, Gomez E, Klunker S, Meyer N, O’Mahony L, Palomares O, Rhyner C, Quaked N, Schaffartzik A, Veen WV, Zeller S, Zimmermann M, Akdis CA, and Switzerland D. Interleukins, from 1 to 37, and interferon-g: Receptors, functions, and roles in diseases. Journal of Allergy and Clinical Immunology. 2011: 127, 701-721. Antar LN, Afroz R, Dictenberg JB, Carroll RC and Bassell GJ. Metabotropic glutamate receptor activation regulates fragile X mental retardation protein and Fmr1 mRNA localization differentially in dendrites and at synapses. Journal of Neuroscience. 2004: 24, 2648-2655. Antonetti DA, Barber AJ, Bronson SK, Freeman WM, Gardner TW and Jefferson LS. Diabetic retinopathy seeing beyond glucose-induced microvascular disease. Diabetes. 2006: 55, 240124011. Bakker CE, Verheij C, Willemsen R, Helm R, Oerlemans F, Vermey M, Bygrave A, Hoogeveen T, and Oostra’t BA, Reyniers E, Boulle KD, D’Hooge R, Cras P, Velzen D, Nagels G, Martin J, Deyn PP. Darby JK, Willems’t PJ. Fmrl knockout mice: a model to study fragile X mental retardation. Cell. 1994: 75, 23-34. Bardoni B, Schenck A and Mandel JL. A novel RNA-binding nuclear protein that interacts with the fragile X mental retardation (FMR1) protein. Human Molecular Genetics. 1999: 8, 25572566. Bassell GJ. and Warren ST. Fragile X syndrome: loss of local mRNA regulation alters synaptic development and function. Cell. 2008: 60, 201-214. Bear MF, Huber KM and Warren ST. The mGluR theory of fragile X mental retardation. Trends in Neurosciences. 2004: 27, 370-377. Bellone C, Luscher C and Mameli M. Mechanisms of synaptic depression triggered by metabotropic glutamate receptors. Cellular and Molecular Life Sciences. 2008: 65, 2913-2923. 22 Boyle L and Kaufmann WE. The behavioral phenotype of FMR1 mutations. American Journal of Medical Genetics. 2010: 145C, 469-676. Brouwer JR, Willemsen R, and Oostra BA. The FMR1 gene and fragile X-associated tremor/ataxia syndrome. American Journal of Medical Genetics. 2009: 105, 782-798. Cai J and Boulton M. The pathogenesis of diabetic retinopathy: old concepts and new questions. Eye. 2002: 16, 242-260. Caldwell RB, Bartoli M, Behzadian MA, El-Remessy AE, Al-Shabrawey M, Platt DH and Caldwell RW. Vascular endothelial growth factor and diabetic retinopathy: pathophysiological mechanisms and treatment perspectives. Diabetes/Metabolism Research and Reviews. 2003: 19, 442-455. Claudia E. Mechanisms of interleukin-1beta release. Immunobiology. 2009: 214, 543-553. Conn J and Pin JP. Pharmacology and functions of metabotropic glutamate receptors. Annual Review of Pharmacology and Toxicology. 1997: 37, 205-237. Contractor A, Klyachko VA and Portera-Cailliau C. Altered neuronal and circuit excitability in fragile X syndrome. Neuron. 2015: 87, 699-715. Coutinho V and Knöpfel T. Metabotropic glutamate receptors: electrical and chemical signaling properties. The Neuroscientist. 2002: 8, 551-561. Crawford DC, Acun JM, and Sherman SL. FMR1 and the fragile X syndrome: Human genome epidemiology review. Genetics in medicine. 2001: 3, 359-371. Curtis TM, Gardiner TA and Stitt AW. Microvascular lesions of diabetic retinopathy: clues towards understanding pathogenesis?. Eye. 2009: 23, 1496-1508. Darnell JC and Klann E. The Translation of Translational Control by FMRP: Therapeutic Targets for Fragile X Syndrome. Nature neuroscience. 2013: 16, 1530-1536. Desai NS, Casimiro TM, Gruber SM, and Vanderklish PW. early postnatal plasticity in neocortex of Fmr1 knockout mice. Journal of Neurophysiology. 2006: 96, 1734-1734. Dhande OS and Huberman AD. Retinal ganglion cell maps in the brain: implications for visual processing. Current Opinion in Neurobiology. 2014: 24, 133-142. Dinarello CA. Proinflammatory Cytokines. Chest. 2000; 118, 503-508. Dinarello CA. Biologic Basis for Interleukin-l in Disease. The Journal of The American Society of Hematology. 1996: 87, 2095-2147. 23 Dinarello CA. Biology of interleukin 1. The FASEB Journal. 1988; 108-123. Ecker JL, Dumitrescu ON, Wong KY, Alam NM, Chen S, LeGates T, Renna JM, Prusky GT, Berson DM and Hattar S. Melanopsin-expressing retinal ganglion-cell photoreceptors: cellular diversity and role in pattern vision. Neuron. 2010: 67, 49-60. Van Essen DC, Anderson CH. Information processing strategies and pathways in the primate retina and visual cortex. In SF Zornetzer, JL Davis, C Lau, eds. Introduction to Neural and Electronic Networks, 2nd ed. Orlando. Academic Press. 1995: 45–76. Ferraguti F and Shigemoto R. Metabotropic glutamate receptors. Cell and Tissue Research. 2006: 326, 383-504. Ferrera VP, Nealey TA and Maunsell JH. Mixed parvocellular and magnocellular geniculate signals in visual area V4. Nature . 1992: 358, 756-758. Finucane B, Abrams L, Cronister A, Archibald AD, Bennett RL and McConkie-Rosell A. Genetic counseling and testing for fmr1 gene mutations: practice guidelines of the national society of genetic counselors. Journal of Genetic Counseling. 2012: 21,752-760. Gabay C, Lamacchia C and Palmer G. IL-1 pathways in inflammation and human diseases. Nature Reviews. Rheumatology. 2010: 6, 232-241. Gallagher A and Hallahan B. Fragile X-associated disorders: a clinical overview. Journal of Neurology. 2012: 295, 401-413. Garber KB , Visootsak J and Warren ST. Fragile X syndrome. European Journal of Human Genetics. 2008: 16, 666-672. Gardiner TA, Archer DB, Curtis T and Stitt AW. Arteriolar involvement in the microvascular lesions of diabetic retinopathy: implications for pathogenesis. Microcirculation. 2007: 14, 25-38. Govindaiah G and Cox CL. Metabotropic glutamate receptors differentially regulate GABAergic inhibition in thalamus. Journal of Neuroscience. 2006: 26, 13443-13453. Grellner W, Georg T and Wilske J. Quantitative analysis of proinflammatory cytokines (IL-1β, IL6, TNF-α) in human skin wounds. Forensic Science International. 2000: 113, 251-264. Hagerman RJ, Berry-Kravis E, Kaufmann WE, Ono MY, Tartaglia N, Lachiewicz A, Kronk R, Delahunty C, Hessl D, Visootsak J, Picker J, Gane L and Tranfaglia M. Advances in the treatment of fragile X syndrome. Pediatrics. 2009. 123:378–390. 24 Harrington ME. The ventral lateral geniculate nucleus and the intergeniculate leaflet: interrelated structures in the visual and circadian systems. Neuroscience and Biobehavioral Reviews. 1997: 21, 705-725. Hagerman RJ, Jackson C, Amiri K; Silverman AC, O’Connor R, and Sobesky W. Girls With Fragile X syndrome: physical and neurocognitive status and outcome. Pediatrics.1992: 89, 395-400. Hendry SH and Reid RC. The koniocellular pathway in primate vision. Annual Review of Neuroscience. 2000: 23, 127-153. Herrero MT, Barcia C and Navarro JM. Functional anatomy of thalamus and basal ganglia. Child's Nervous System Journal. 2002: 18, 386-404. Huberman AD and Niell CM. What can mice tell us about how vision works?. Trends in Neurosciences. 2011: 34, 465-473. Huberman AD, Wei W, Elstrott J, Stafford BK, Feller MB, and Barres BA. Genetic identification of an on-off direction- selective retinal ganglion cell subtype reveals a layer-specific subcortical map of posterior motion. Neuron. 2009: 62, 327-334. Huddleston LB, Visootsak J and Sherman SL. Cognitive aspects of fragile X syndrome. Wiley interdisciplinary reviews. Cognitive Science. 2014: 5, 501-508. Irvin GE, Casagrande VA, and Norton TT. Center/surround relationships of magnocellular, parvocellular, and koniocellular relay cells in primate lateral geniculate nucleus. Visual Neuroscience. 1993: 10, 363-373. Irwin SA, Idupulapati M, Gilbert ME, Harris JB, Chakravarti AB, Rogers EJ, Crisostomo RA, Larsen BP, Mehta A, Alcantara CJ, Patel B, Swain RA, Weiler IJ, Oostra BA and Greenough WT. Dendritic spine and dendritic field characteristics of layer V pyramidal neurons in the visual cortex of fragile-X knockout mice. American Journal of Medical Genetics. 2002: 111, 140-146. Jeffrey CP and Pin J. Pharmacology and functions of metabotropic glutamate receptors. Annual Review of Pharmacology and Toxicology. 1997: 37, 205-237. Joussen AM, Poulaki V, Le ML, Koizumi K, Esser C, Janicki H, Schraermeyer U, Kociok N, Fauser S, Kirchhof B, Kern TS, and Adamis AP. A central role for inflammation in the pathogenesis of diabetic retinopathy. The FASEB Journal. 2004: 18, 1450-1468. Kempen JH, O'Colmain BJ, Leske MC, Haffner SM, Klein R, Moss SE, Taylor HR, Hamman RF, West SK, Wang JJ, Congdon NG, and Friedman DS. The prevalence of diabetic retinopathy among adults in the united states. The Eye Diseases Prevalence Research Group. Archives of ophthalmology. 2004: 122, 552-563. 25 Kim M, Bellini M and Ceman S. Fragile X mental retardation protein FMRP binds mRNAs in the nucleus. Molecular and Cellular Biology. 2009: 29,214-228. Kooy R F. Of mice and the fragile X syndrome. Trends in genetics. 2003: 19, 148-154. Kowluru RA and Odenbach S. Role of interleukin-1b in the pathogenesis of diabetic retinopathy. British journal of ophthalmology. 2004: 88, 1343-1347. Krahe TE, El-Danaf RN, Dilger EK, Henderson SC, and Guido W. Morphologically distinct classes of relay cells exhibit regional preferences in the dorsal lateral geniculate nucleus of the mouse. Journal of Neuroscience. 2011: 31, 17437–17448. Kroll P, Rodrigues EB and Hoerle S. Pathogenesis and classification of proliferative diabetic vitreoretinopathy. Ophthalmologica Journal. 2007: 22, 78-94. Larson, J., Jessen, R.E., Kim, D., Fine, A.K., and du Hoffmann, J. (2005). Agedependent and selective impairment of long-term potentiation in the anterior piriform cortex of mice lacking the fragile X mental retardation protein. Journal of Neuroscience. 25, 9460–9469. Leist M, Datunashvilli M, Kanyshkova T, Zobeiri M, AissaouiA, Cerina M, Romanelli MN, Pape H and Budde T. Two types of interneurons in the mouse lateral geniculate nucleus are characterized by different h-current density. Scientific Report. 2016: 6, 24904-24919. Li J, Pelletier MR, Velazquez JP and Carlen PL. Reduced cortical synaptic plasticity and glur1 expression associated with fragile x mental retardation protein deficiency. Molecular and Cellular Neurosciences. 2002: 19, 138-151. Li P, Allen H, Banerjee S, Franklin S, Herzog L, Johnston C, McDowell J, Paskind M, Rodman L, Salfeld J, Towne E, Tracey D, Wardwell S, Wei F, Wong W, Kamen R, and Seshadri T. Mice deficient in il-lp-converting enzyme are defective in production of mature il-lp and resistant to endotoxic shock. Cell. 1995: 80, 401-411. Maunsell J, Ghose GM, Assad JA, Mcadams CJ, Boudreau CE, and Noerager BD. Visual response latencies of magnocellular and parvocellular LGN neurons in macaque monkeys. Visual Neuroscience. 1999: 16, 1-14. Martín CA, Crespo M, and Cailliau CP. Delayed stabilization of dendritic spines in fragile X mice. Journal of Neuroscience. 2010, 30:7793–7803. Martin JP and Bell J. A pedigree of mental defect showing sex-linkage. Journal of Neurology and Psychiatry. 1943: 6, 154-157. 26 Meldrum B and Garthwaite J Excitatory amino acid neurotoxicity and neurodegenerative disease. Trends in Pharmacological Sciences. 1990: 11, 379-387. Mizel SB. The interleukins. The FASEB Journal. 1989: 3, 2379-2390. Monaghan DT, Bridges RJ and Cotman CW. The excitatory amino acid receptors: their classes, pharmacology, and distinct properties in the function of the central nervous system. Annual Review of Pharmacology and Toxicology. 1989: 29, 365-402. Morgan TH. An attempt to analyze the constitution of the chromosomes on the basis of sexlimited inheritance in Drosophila. Journal of Experimental Zoology. 1911: 11, 365-413. Morin LP and Blanchard JH. Descending projections of the hamster intergeniculate leaflet: Relationship to the sleep/arousal and visuomotor systems. Journal of Comparative Neurology. 2005: 487, 204-216. Nakajima Y , Iwakabe H, Akazawa C, Nawa H, Shigemoto R, Mizuno N and Nakanishi S . Molecular characterization of a novel retinal metabotropic glutamate receptor mGluR6 with a high agonist selectivity for L-2-amino-4-phosphonobutyrate. Journal of Biological Chemistry. 1993: 268, 11868-11873. Berlucchi G and Buchtel HA. Neuronal plasticity: historical roots and evolution of meaning. Experimental Brain Research. 2009: 192, 307-319. Niswender CM and Conn PJ. Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annual Review of Pharmacology and Toxicology J. 2010: 50,295-322. Parnet P, Kelley K W, Bluthé R and Dantzer R. Expression and regulation of interleukin-1 receptors in the brain. Role in cytokines-induced sickness behavior. Journal of Neuroimmunology. 2002: 125, 5-14. Engerman RL. Pathogenesis of diabetic retinopathy. Diabetes. 1989: 38, 1203-1206. Pengarikano O, Mulle LG and Warren ST. The pathophysiology of fragile X syndrome. Annual Review of Genomics and Human Genetics. 2007: 8; 109-129. Pfeiffer BE and Huber KM. The state of synapses in fragile x syndrome. Neuroscientist. 2009: 15, 549-567. Pinault D. The thalamic reticular nucleus: structure, function and concept. Brain research reviews. 2004: 46, 1-31. Qiu A, Bian Z, Mao P and Liu Q. IL-17A exacerbates diabetic retinopathy by impairing. Müller cell function via Act1 signaling. Experimental & Molecular Medicine. 2016: 48, 280-291. 27 Rafols JA and Valverde F. The structure of the dorsal lateral geniculate nucleus in the mouse. a golgi and electron microscopic study. Journal of Comparative Neurology. 2004: 150, 303-331. Ridaura-Ruiz L, Quinteros-Borgarello M, Berini-Aytés L and Gay-Escoda C. Fragile X-syndrome: literature review and report of two cases. Med Oral Patol Oral Cir Buccal. 2009: 1, 434-439. Riedel G, Platt B, and Micheau J. Glutamate receptor function in learning and memory. Behavioral Brain Research. 2003: 140, 1-47. Rioult-Pedotti M, Friedman D and Donoghue JP. Learning-induced LTP in neocortex. Science. 2000: 290, 533-536. Rothwell NJ and Luheshi GN. Interleukin 1 in the brain: biology, pathology and therapeutic target. Trends in Neurosciences. 2000: 23, 678-625. Rousseau F , Heitz D , Tarleton J, MacPherson J , Malmgren H , Dahl N, Barnicoat A, Mathew C, Mornet E, Tejada I, Maddalena A, Spiegel R, Schinzel A, Marcos JA, Schorderet DF, Schaap T, Maccioni L, Russo S, Jacobs PA, Schwartz C, and Mandel JL. A multicenter study on genotypephenotype correlations in the fragile x syndrome, using direct diagnosis with probe stb12.3: the first 2,253 cases. American Journal of Human Genetics. 1994: 55, 225-237. Sabaratnam M. Fragile-X syndrome. Psychiatry. 2006: 5, 325-330. Schenck A, Bardoni B, Moro A, Bagni C and Mandel J. A highly conserved protein family interacting with the fragile X mental retardation protein (FMRP) and displaying selective interactions with FMRP-related proteins FXR1P and FXR2P. PNAS. 2001: 98, 8844-8849. Sullivan K, Hatton D, Hammer J, Sideris J, Hooper S, Ornstein P and Bailey D. ADHD symptoms in children with FXS. American Journal of Medical Genetics. 2006: 140, 2275-2288. Swenson R. Review of clinical and functional neuroscience. Educational Review Manual in Neurology 2006. Sherman SM. Functional organization of the W-, X-, and Y-cell pathways: a review and hypothesis. In: Progress in Psychobiology and Physiological Psychology. 1985: 233–314. New York: Academic. Sherman SM. Tonic and burst firing: dual modes of thalamocortical relay. Trends in Neurosciences. 2001: 26, 122-126. Sherman SM and Guillery RW. The role of the thalamus in the flow of information to the cortex. The Royal Society of London. 2002: 357, 1695-1708. 28 Sheridan SD, Theriault KM, Reis SA, Zhou F, Madison JM, Daheron L, Loring JF and Haggarty SJ. Epigenetic characterization of the FMR1 gene and aberrant neurodevelopment in human induced pluripotent stem cell models of fragile X syndrome. PLOS One. 2011; 6, e26203. Sidorov MS, Auerbach BD and Bear MF. Fragile X mental retardation protein and synaptic plasticity. Molecular Brain. 2013: 6, 15. Solomon SG, White AJ and Martin PR. Extraclassical receptive field properties of parvocellular, magnocellular, and koniocellular cells in the primate lateral geniculate nucleus. Journal of Neuroscience. 2002: 22, 338-350. Sur M, Garraghty PE and Roe AW. experimentally induced visual projections into auditory thalamus and cortex. Science. 1988: 242, 1437-1441. Tanabe Y, Masu M, Ishii T, Shigemoto R and Nakanishi S. A family of metabotropic glutamate receptors. Neuron. 1992: 8, 169-179. Tang J and Kern TS. Inflammation in diabetic retinopathy. Progress in Retinal and Eye Research. 2011: 30, 343-358. Tang J, Jimenez SCA, Chakraborty S and Schultz SR. Visual receptive field properties of neurons in the mouse lateral geniculate nucleus. PLOS One. 2016: 11, e0146017- e0146051. Turner G, Robinson H, Laing S and Purvis-Smith S. Preventive screening for the fragile X syndrome. The New England journal of medicine. 1986: 315, 607–609. Vujosevic S and Midena E. Retinal layers changes in human preclinical and early clinical diabetic retinopathy support early retinal neuronal and Müller cells alterations. Journal of Diabetes Research. 2013: 2013, 1-8. Ward LM. The thalamus: gateway to the mind. Wiley interdisciplinary reviews. Cognitive science. 2013: 4, 609-622. Weber A, Wasiliew P, and Kracht M. Interleukin-1 (IL-1) pathway. Science Signaling. 2010: 3, 105 (1-6). Weliky M and Katz LC. Correlational structure of spontaneous neuronal activity in the developing lateral geniculate nucleus in vivo. Science. 1999: 285, 599-604. Whitmire W, Al-Gayyar MH, Abdelsaid M, Yousufzai BK and El-Remessy AB. Alteration of growth factors and neuronal death in diabetic retinopathy: what we have learned so far. Molecular Vision. 2011: 17, 300-308. 29 Wilson BM and Cox CL. Absence of metabotropic glutamate receptor-mediated plasticity in the neocortex of fragile X mice. PNAS. 2007: 104, 2454-2459. Wisden W and Seeburg PH. Mammalian ionotropic glutamate receptors. Current Opinion in Neurobiology. 1993: 3, 291-298. Wong M, Bongiorno PB, Rettori V, McCann SM, and Licinio J. Interleukin (IL) 1β, IL-1 receptor antagonist, IL-10, and IL-13 gene expression in the central nervous system and anterior pituitary during systemic inflammation: Pathophysiological implications. PNAS. 1997: 94, 227-232. Zhang, HQ, Murray GM, Coleman GT, Turman AB, Zhang SP, and Rowe MJ. Evidence for two complementary patterns of thalamic input to the rat somatosensory cortex. Brain Research. 1988: 463, 346-351. Zhang J, Liu R, Kuang H, Gao X and Liu H. Protective treatments and their target retinal ganglion cells in diabetic Retinopathy. Brain Research Bulletin. 2017: 132, 53-60. Zhang L, Liang Z, Zhu P, Li M, Yi Y, Liao W and Su T. Altered intrinsic properties and bursting activities of neurons in layer IV of somatosensory cortex from Fmr-1 knockout mice. Experimental Neurology. 2016: 280, 60-69. Zhao MG, Toyoda H, Ko SW, Ding HK, Wu L and Zhuo M. Deficits in trace fear memory and longterm potentiation in a mouse model for fragile x syndrome. Journal of Neuroscience. 2005: 25, 7358-7392. 30 Chapter 2 Postsynaptic actions of mGluR activation in neocortical neurons are unaltered in Fmr1 KO mice Abstract Fragile X Syndrome (FXS) is one of the most common forms of inherited mental retardation. In FXS, there is a mutation in the fragile X mental retardation 1 (Fmr1) gene resulting in a lack of FMRP production, which serves as an important element in many neuronal processes. Previous studies have indicated that FXS patients as well as an animal model of FXS have altered neuroanatomical characteristics. FXS is associated with neuronal defects; previous studies have also found alterations in synaptic plasticity that is dependent on the activation of metabotropic glutamate receptors in the Fmr1 KO mice, a fragile X mice model. In the neocortex, mGluR-dependent long-term potentiation is significantly attenuated in Fmr1 KO animals. In this study, we determined if mGluR-dependent regulation of neuronal excitability is also altered in Fmr1 KO animals. Changes in mGluR-dependent functions could serve as potential mechanisms for the dampening of cognitive abilities associated with FXS. We tested the effects of selective mGluR agonists on the excitability of deep layer neocortical pyramidal neurons in Fmr1 KO and wild type (WT) mice. We found that postsynaptic actions of mGluR activation on membrane potentials were unaltered in Fmr1 KO mice, leading to the conclusion that short term membrane potential changes produced by mGluR activation are probably not involved in the neurological conditions observed in FXS. 31 Introduction Fragile X Syndrome (FXS) is the most common form of inherited mental retardation. This condition results from the loss of fragile X mental retardation protein (FMRP; Song et al., 2003). In addition to cognitive deficits, FXS patients may display a wide range of other conditions including anxiety, compulsive behaviors, epilepsy, autism spectrum disorder, and poor motor coordination (Bear et al., 2004). The current pharmacological treatment for FXS patients is symptomatic, with mixed results, which suggest that there are likely multiple mechanisms that underlie the neurological alterations. This is further supported in that the magnitude of cognitive attenuation varies from mild to severe conditions. Alterations in metabotropic glutamate receptor (mGluR) function appear to play a crucial role in the defects associated with FXS. In Fmr1 KO mice, group I mGluRs are overactivated, due to the absence of FMRP; this absence will lead to the increase of mRNA translation (Chuang et al., 2005; Huber et al., 2002). In the hippocampus, Fmr1 KO mice showed that protein synthesis occurs after group I mGluR activation, leading to the enhancement of long term depression (LTD; Bear et al., 2004). It has been also hypothesized that FMRP, which is found in dendritic spines after being encoded by the Fmr1 gene, has a major influence in slowing down local protein syntheses at synapses as a response of metabotropic glutamate receptors activation (Bassell and Gross, 2008; Weiler and Greenough, 1999). 32 An early finding in both FXS patients and Fmr1 KO mice was an alteration of dendritic spine density and morphology in the neocortex and hippocampus. There is an increased proportion of immature-appearing spines within the FXS condition, and thus led to the speculation that these abnormalities may be related to the cognitive dampening in FXS (Comeryet al., 1997; Levenga et al., 2011). While the functional consequence of these immature spines is unclear, these changes could impact synaptic connectivity and subsequent synaptic efficacy. Multiple forms of synaptic plasticity in Fmr1 KO mice have been broadly studied in many brain regions. In neocortex, long term potentiation (LTP) that is dependent on mGluR5 activation is significantly attenuated in deep layers of the Fmr1 KO mice (Wilson and Cox, 2007). Similarly, mGluR-dependent LTP in the lateral amygdala is impaired in the Fmr1 KO mice (Suvrathana et al., 2010). In addition to alterations in synaptic plasticity, certain intrinsic properties of neurons are altered in Fmr1 KO mice. The Fmr1 KO mice excitatory neurons exhibit more neuronal excitability and this excitement is driven by the excessive activation mediated by group I mGluR (Hays et al., 2011). In pyramidal neurons from somatosensory cortex, Zhang et al. (2016) found that Fmr1 KO mice have an increased neuronal excitability, which can be seen in the abnormally higher action potential firing frequency when compared with a WT population. It is hypothesized that cognitive and behavioral defects associated with FXS is due to changes in neocortical excitability and functions (Gibson et al., 2008; Till et al., 2012). These increases in intrinsic excitability of excitatory neurons tend to push the system to a hyperexcitable state which could account for the circuit hyperexcitability observed in the neocortical excitatory neurons (Pfeiffer and Huber, 2009). It was also reported by Gibson et al., 2008 that the increase 33 in excitability in Fmr1 KO mice might be due to the higher input resistance and reduced membrane capacitance of layer 4 excitatory somatosensory cortical neurons. The synaptic plasticity alteration as well as neuronal hyperexcitability associated with FXS are significant elements behind the phenotypes seen in FXS. The activation of mGluRs has been associated with many brain functions including learning and memory, anxiety, pain, epilepsy, and cognitive development (Niswender and Conn, 2010). Clearly there appears to be differential effects on mGluR-dependent functions in the Fmr1 KO animals. These finding indicate that the lack of FMRP will result in synaptic plasticity alterations, and it is suggested that these modifications could be region specific (Hays et al., 2011; Gibson et al., 2008; Wilson and Cox, 2007). Considering earlier work from our laboratory, where we found a decreased mGluRmediated plasticity in the neocortex (Wilson and Cox, 2007), we wanted to follow up on this finding and determine if the neuronal responsiveness to mGluR activation is altered in Fmr1 KO mice. We previously found that mGluR-dependent LTP is reduced in Fmr1 KO, and predict that there may be decreased levels of group I mGluRs in the neocortex, and thus postsynaptic responses to mGluR activation may be attenuated in Fmr1 KO mice. In this experiment, we will determine if selective mGluR agonists differentially alter membrane potential of deep layer (V/VI) pyramidal neurons from WT and Fmr1 KO mice. 34 Materials and Methods All animal care and experimental procedures used in this study were approved by the Michigan State University Institutional Animal Care and Use Committee. Both Fmr1 KO and WT mice used in these experiments were maintained on a C57Bl/6J background (Bakker et al., 1994). Mice of either sex, ages 16 to 21 days, were deeply anaesthetized with 2-4% isoflurane. The animal was decapitated, and the brain was quickly removed and placed in a cold (<4 °C), oxygenated slicing solution containing (in mM): 2.5 KCl, 1.25 NaH2PO4, 10.0 MgSO4, 0.5 CaCl2, 234.0 sucrose, 10.0 glucose, and 26.0 NaHCO3. Slices (300 μm thick) were cut on a vibrating tissue slicer in the coronal plane, and then incubated in a heated holding chamber (35 °C) containing oxygenated physiological saline containing (in mM): 26.0 NaHCO3, 2.5 KCl, 10.0 glucose, 126.0 NaCl, 1.25 NaH2PO4, 2.0 MgCl2, and 2.0 CaCl2 for at least 30 minutes and then maintained at room temperature until used. Individual slices were then moved to a submersion recording chamber, and continuously super fused with oxygenated physiological saline (2-3 ml/min) and maintained at 32 ± 1 °C. Whole cell recordings were obtained from layer V/VI pyramidal neurons in somatosensory cortex. Pipettes had tip resistances of 2-5 MΩ when filled with internal solution containing (in mM): 117.0 K-gluconate, 13.0 KCl, 1.0 MgCl2, 0.07 CaCl2, 0.1 EGTA, and 10.0 HEPES, with a pH of 7.30 and an osmolarity of 290. Data were acquired using Multiclamp 700B amplifier, signals were digitized (10 kHz) and filtered (10 kHz) and subsequent analyses were done using pClamp software (Molecular Devices, Sunnyvale, CA). The liquid junction potential (10 mV) was corrected for in all recordings. 35 Agonists were bath applied by infusion into the bath via a syringe pump (Paul and Cox, 2010). The following selective mGluR agonists were used: group I mGluR: (S)-3,5dihydroxyphenylglycine hydrate (DHPG) and group II mGluR: 4-amino-2,4- pyrrolidinedicarboxylic acid (APDC). The sodium channel blocker tetrodotoxin (TTX) was added to the bath prior to mGluR agonist application to prevent action potentials in both pre- and post-synaptic neurons. All chemicals were purchased from Tocris (St Louis, MO, USA) and Sigma (St. Louis, MO, USA). All data statistical analysis were done using ANOVA test, where the statistical significance was p<0.05. All error bars in figures represent the SME. 36 Results Recordings were obtained from 30 WT and 23 Fmr1 KO layer V/VI pyramidal neurons of somatosensory cortex. In WT neurons, the average resting membrane potential (RMP) was 72.1 ± 3.9 mV (n=30) and apparent input resistance (Rin) averaged 266.4 ± 141.8 MΩ (n=30). In neurons from Fmr1 KO animals, the resting membrane potential averaged -74.4 ± 4.6 mV (n=23) and the apparent input resistance averaged 384.9 ± 112.0 MΩ (n=23). The membrane potential did not significantly differ from between WT and Fmr1 KO neurons (RMP: p=0.06, ttest); however, the input resistance of neurons from Fmr1 KO animals was significantly greater that those from WT animals (Rin: p= 0.001, t-test). RMP (mV) Rin (MΩ) mice. WT Fmr1 KO -72.1 ± 3.9 -74.4 ± 4.6 p= 0.060, (n=30) (n=23) t-test 266.4 ± 141.8 384.9 ± 112.0 *p= 0.001, (n=30) (n=23) t-test Table 2.1. Intrinsic properties of layer V/VI pyramidal neurons from WT and Fmr1 KO 37 Group I mGluR We tested the effect of the selective group I mGluR agonist DHPG on the membrane potential of somatosensory pyramidal neurons. After establishing a stable baseline, DHPG was briefly bath applied for 60 seconds. At the lowest concentration tested, DHPG (5 μM) produced a small depolarization in WT neurons that averaged 0.6 ± 0.2 mV (n=9). At higher concentrations, DHPG produced a larger membrane depolarization (Figure 2.1, 25 μM: 3.4 ± 2.1 mV, n=17; 100 μM: 4.2 ± 2.7 mV, n=15). In neurons from Fmr1 KO animals, DHPG also produced a dose dependent increase in membrane potential depolarization (Figure 2.1, 5 μM: 0.5 ± 0.2 mV, n=8; 25 μM: 3.3 ± 1.8 mV, n=15; 100 μM: 3.6 ± 2.1 mV, n=15). The magnitude of the membrane depolarization produced by DHPG did not differ between WT and Fmr1 KO neurons at each concentration tested, (p=0.062, ANOVA test). 38 Figure 2.1. Group I mGluR agonist DHPG produces similar membrane depolarizations in pyramidal neurons from WT and Fmr1 KO animals. A. Voltage recordings from representative WT and Fmr1 KO neurons. DHPG produced a concentration dependent increase in amplitude in both types of neurons. B. Summary of population data (mean ± SEM). 39 Group II mGluR In WT neurons, the group II mGluR agonist APDC (10 μM), produced a small membrane hyperpolarization (Figure 2.2, 1.0 ± 0.5 mV, n=6). At higher concentrations, the hyperpolarization was larger (75 μM: 1.9 ± 0.8 mV, n=11; 150 μM: 1.7 ± 0.9 mV, n=13). In neurons from Fmr1 KO animals, APDC produced a dose dependent increase in membrane hyperpolarization (10 μM: 0.6 ± 0.4 mV, n=8; 75 μM: 1.2 ± 0.5 mV, n=8; 150 μM: 1.4 ± 0.9 mV, n=13). The membrane depolarization produced by APDC did not differ between WT and Fmr1 KO neurons at each concentration tested, (p= 0.24, ANOVA test). 40 Figure 2.2. Group II mGluR agonist APDC produces similar membrane hyperpolarizations in pyramidal neurons from WT and Fmr1 KO animals. A. Voltage recordings from representative WT and Fmr1 KO neurons. APDC produced a concentration dependent increase in amplitude. B. Summary of population data (mean ± SEM). 41 Discussion In this study, we found that the postsynaptic membrane changes produced by activation of either group I or II mGluRs in layer V/VI pyramidal neurons did not differ between WT and Fmr1 KO mice, indicating that the loss of FMRP is not affecting the membrane potential changes produced by short term activation of type I and II mGluRs. Previous studies suggest that stimulation of mGluR can lead to increases in local protein synthesis (Bear and Huber, 2004; Dölen and Bear, 2008), and in Fmr1 KO, this protein synthesis is dampened (Bear et al., 2004). Numerous studies report that protein synthesis can occur within dendritic spines (Steward and Schuman, 2001). Such intracellular alterations by mGluR activation do not seem to impact the membrane potential changes. In this work, the lack of quantitative differences between Fmr1 KO and WT mice when tested with selective group I and II mGluR agonists indicates that the defects associated with mGluR activation are primarily associated with long term plasticity (Zhang and Alger, 2010). We have previously shown that LTP in deep layer neocortex is strongly reduced in Fmr1 KO mice, and this reduction is due to a decrease in mGluR5-dependent plasticity (Wilson and Cox, 2007). From that study, we concluded that there was a reduction in mGluR5-mediated activity in the Fmr1 KO mice compared with the WT mice. Our current results show no significant differences between the membrane response to group I and II mGluR activation in WT and Fmr1 KO mice. We would predict that the membrane depolarization depends on mGluR1 activation (Martín et al., 2010; Desai et al., 2006), or alternatively is likely due to activation of mGluR1 , and that mGluR5 activation is tied to LTP induction in the deep layer 42 neocortex. In this study, we show that even with the lack of FMRP in the Fmr1 KO mice, layer V/VI somatosensory neurons are still able to produce an intact short term membrane potential changes, short term depolarization or short term hyperpolarization. It might be that the binding affinity of the mGluRs is affected in the absence of FMRP. When the plasticity was measured in our previous study, the Fmr1 KO mice potentiated only for 20 minutes, while the WT mice showed a potentiation to up to 60 minutes after tetanic stimulation in the neocortex. In this study, the lack of alteration could be due to intact mGluR function in Fmr1 KO mice for that limited time, and when this time is exceeded, their function will start to deteriorate, which could be directly linked to the mGluR binding ability alteration in the Fmr1 KO. If this is the case with Fmr1 KO mice, we could pharmacologically increase the mGluR binding ability and restore its function. 43 REFERENCES 44 REFERENCES Antar LN, Afroz R, Dictenberg JB, Carroll RC and Bassell GJ. Metabotropic glutamate receptor activation regulates fragile X mental retardation protein and Fmr1 mRNA localization differentially in dendrites and at synapses. Journal of Neuroscience. 2004, 24:2648-2655. Bakker CE, Verheij C, Willemsen R, Helm R, Oerlemans F, Vermey M, Bygrave A, Hoogeveen T, and Oostra’t BA, Reyniers E, Boulle KD, D’Hooge R, Cras P, Velzen D, Nagels G, Martin J, Deyn PP. Darby JK, Willems’t PJ. Fmrl Knockout Mice: A model to study fragile x mental retardation. Cell. 1994, 75:23-34. Bassell JG and Gross C. Reducing glutamate signaling pays off in fragile X. Nature Medicine. 2008, 14:249-250. Bear MG, Huber KM, and Warren ST. The mGluR theory of fragile X mental retardation. Trends in Neurosciences. 2004, 27:370-377. Chuang S, Zhao W, Bauchwitz R, Yan Q, Bianchi R, and Wong RKS. Prolonged epileptiform discharges induced by altered group I metabotropic glutamate receptor-mediated synaptic responses in hippocampal slices of a fragile X mouse model. Journal of Neuroscience. 2005, 25:8048 – 8055. Comery TA, Harris JB, Willems PJ, Oostra BA, Irwin SA, Weiler IJ and Greenough WT. Abnormal dendritic spines in fragile X knockout mice: maturation and pruning deficits. PNAS. 1997: 94, 5401-5404. Desai SN, Casimiro MT, Gruber MS, and Vanderklish WP. Early postnatal plasticity in neocortex of Fmr1 knockout mice. Journal of Neurophysiology. 2006, 96:1734 –1745. Dölen G and Bear MG. Role for metabotropic glutamate receptor 5 (mGluR5) in the pathogenesis of fragile X syndrome. Journal of Physiology. 2008, 586:1503-1508. Gibson RJ, Bartley FA, Hays AS, and Huber KM. Imbalance of neocortical excitation and inhibition and altered up states reflect network hyperexcitability in the mouse model of fragile X syndrome. Journal of Neurophysiology. 2008, 100:2615–2626. Hays SA, Huber KM. and Gibson JR. Altered neocortical rhythmic activity states in Fmr1 KO mice are due to enhanced mGluR5 signaling and involve changes in excitatory circuitry. Neurobiology of Disease. 2011, 31:14223–14234. 45 Huber MK, Gallagher MS, Warren TS, and Bear MG. Altered synaptic plasticity in a mouse model of fragile X mental retardation. Proceedings of the National Academy of Sciences. 2002, 99:7746-7750. Levenga J, Vrij FMS, Buijsen R, Li T, Nieuwenhuizen MI, Pop A, Oostra AB, and Willemsen R. Subregion-specific dendritic spine abnormalities in the hippocampus of Fmr1 KO mice. Neurobiology of Learning and Memory. 2011, 95:467–472 Martín CA, Crespo M, and Cailliau CP. Delayed stabilization of dendritic spines in fragile X mice. Journal of Neuroscience. 2010, 30:7793–7803. Merlin LR. Differential roles for mGluR1 and mGluR5 in the persistent prolongation of epileptiform bursts. Journal of Neurophysiology. 2001, 87:621–625. Mountcastle BV. The columnar organization of the neocortex. Brain. 1997, 120:701–722. Niswender CM and Conn PJ. Metabotropic glutamate receptors: physiology, pharmacology, and disease. Annual Review of Pharmacology and Toxicology. 2010, 50:295-322. Paul K and Cox CL. Excitatory actions of substance P in the rat lateral posterior nucleus. European Journal of Neuroscience. 2010, 31:1-13. Pfeiffer BE and Huber KM. The state of synapses in fragile X syndrome. The Neuroscientist. 2009, 15:549-567. Ronesi AJ and Huber MK. Metabotropic glutamate receptors and fragile X mental retardation protein: partners in translational regulation at the synapse. Science Signaling. 2008, 1:1-6. Song FJ, Barton P, Sleightholme V, Yao GL, and Fry-Smith A. Screening for fragile X syndrome: a literature review and modelling study. Health Technology Assessment. 2003; 7, 16-20. Steward O and Schuman ME. Protein synthesis at synaptic sites on dendrites. Annual Review of Neuroscience. 2001, 24:299-325. Suvrathana A, Hoefferb CA, Wongb H, Klannb E, and Chattarjia S. Characterization and reversal of synaptic defects in the amygdala in a mouse model of fragile X syndrome. PANS. 2010, 107:11591-11596. Till MS, Wijetunge SL, Seidel GV, Harlow E, Wright KA, Bagni C, Contractor A, Gillingwater HT, and Kind CP. Altered maturation of the primary somatosensory cortex in a mouse model of fragile X syndrome. Human Molecular Genetics. 2012, 21:2143-2156. 46 Tsiouris A J and Brown WT. Neuropsychiatric symptoms of fragile X syndrome: pathophysiology and pharmacotherapy. CNS Drugs. 2004, 18:687-703. Weiler IJ and Greenough WT. Synaptic synthesis of the fragile X protein: possible involvement in synapse maturation and elimination. American Journal of Medical Genetics. 1999, 83:248–252. Wilson BM and Cox CL. Absence of metabotropic glutamate receptor-mediated plasticity in the neocortex of fragile X mice. PNAS. 2007, 104:2454-2459. Zhang L and Alger BE. Enhanced endocannabinoid signaling elevates neuronal excitability in Fragile X syndrome. Journal of Neuroscience. 2010, 30:5724–5729. Zhang L, Liang Z, Zhu P, Li M, Yi Y, Liao W, and Su T. Altered intrinsic properties and bursting activities of neurons in layer IV of somatosensory cortex from Fmr-1 knockout mice. Experimental Neurology. 2016, 280:60-69. 47 Chapter 3 Intraocular injection of Interleukin 1-β alters intrinsic properties of dorsal lateral geniculate nucleus relay neurons Abstract Diabetic retinopathy is the most common diabetic related condition and it is considered to be the leading cause of blindness in working age adults. The circulation of hyperglycemic blood in retinal blood vessels lead to swelling of the vessels and eventually leakage of blood content, which will activate Muller as well as microglia cell. This activation will lead to the initiation of inflammatory response, which in turn will lead to the activation of many proinflammatory cytokines, including interleukin 1-β (IL-1β). The activation of Muller and microglia cells will also lead to the excess production of glutamate, which can directly lead to the over-excitation of retinal ganglion cells (RGCs) leading to their death. After mimicking the inflammatory response associated with diabetic retinopathy, we show that the IL-1β injection alters the lateral geniculate nucleus relay neurons. After inducing inflammation by injecting a pro-inflammatory cytokine IL-1β into the intraocular fluid, we studied its effect on the dorsal lateral geniculate nucleus (dLGN) relay neurons and we found that after 72 hours post IL-1β injection, the resting membrane potential was hyperpolarized, indicating more inhibition of relay neurons. Further, two weeks after injection, the average maximum firing rate was higher. It was also noted that dLGN relay neurons did not show any changes directly after the injection, indicating that the inflammatory response led to retinal cells alteration, which in turn influenced dLGN neurons. 48 Introduction Diabetic retinopathy is a common diabetes related complication, and it is the leading cause of blindness in US adults (Williams et al., 2004). When hyperglycemic blood circulates in the retina, it causes abnormalities in blood vessels, including the presence of nonperfused capillaries and thickening of the basement membrane (Vincent and Mohr, 2007), which in turn produces an inflammatory response leading to early stages of diabetic retinopathy (Sheetz and King 2002; Kohner et al., 1995). There are two main stages in the development of diabetic retinopathy: non-proliferative diabetic retinopathy (NPDR) and proliferative diabetic retinopathy (PDR). In the NPDR, the initiation of inflammatory cascades caused by the hyperglycemic blood will lead to blood vessels leak and cause retina swelling; this initiation of the inflammatory response will be associated with the activation of caspase 1, glia and microglia cells. The activation of caspase 1 will lead to the activation of proinflammatory cytokines as well as the acceleration in apoptotic cell death (Vincent and Mohr, 2007). On the other hand, the activation of glia and microglia cells will lead to the production of glutamate; this glutamate production will eventually lead to RGC excitotoxicity (Vincent and Mohr, 2007; Kempen et al., 2004; Penn et al., 2008). The advance stage of diabetic retinopathy PDR, in which new blood vessels grow within the retina, and these new vessels are more fragile and can rupture more easily leading to chronic inflammation (Penn et al., 2008). Chronic inflammation is usually associated with an increase in vascular permeability, inflammatory cell infiltration, edema, tissue destruction, and neovascularization, which most can be seen in diabetic retinopathy, and these inflammatory responses further maintain the inflammatory cycle as a positive feedback (Adamis and Adrienne, 2008). This advanced stage of diabetic retinopathy 49 can lead to dark spot vision or can even result in complete blindness when the edema in the blood vessels increase to a level where no blood supply reaches the retina, causing an ischemic retina (Vincent and Mohr, 2007; Ciulla et al., 2003; Kempen et al., 2004). IL-1β is a pro-inflammatory cytokine, which is expressed in many cells in the immune system (Eskan et al., 2008). Studies have shown that IL-1β expression is increased during inflammatory responses, such as diabetic retinopathy (Michelle and Akpek, 2014; Antonetti et al., 2006). Many studies have reported that diabetic patients tend to have a higher rate of RGC apoptosis; this was also seen in a diabetic mouse model (Kern and Barber, 2008). It was also reported that after the initiation of the inflammatory cascade, the activation of inflammatory factors, including IL-1β, will lead to the thinning of the inner plexiform layer, where RGC dendrites are located, causing the loss of RGCs function and contributes in their death. (Kern and Barber, 2008; Lieth et al., 2000). The dLGN is the visual relay nucleus of the thalamus, where sensory input is received by the retina and sent to the visual cortex (Wong, 1999). In mice, RGCs show structural changes, including swelling of the axons after 3 months of diabetes (Kern and Barber, 2008). Alterations in RGC function could directly affect dLGN neurons. During diabetic retinopathy, RGCs and glial cells undergo many inflammatory responses including increases in inflammatory factors, which would enhance apoptosis (Gardner et al., 2002; Barber et al., 1998). In mice, the peak of RGC apoptosis will occur 22 weeks after hyperglycemia (Kern and Barber, 2008). Cell death in the retina can lead to elevated levels of glutamate, which can lead to hyperexcitability or excitotoxicity of the RGCs (Barber et al., 1998). 50 In general, the pathogenesis of diabetic retinopathy is chronic, occurs over long period of time, months or even years. Here we hypothesize that by the intraocular injection of IL-1β, we will be able to induce an acute inflammatory response, which in turn allows us to test its downstream effect on the dLGN in shorter time scale. In this study, we will determine whether an inflammatory response after intraocular injections of IL-1β leads to a downstream alteration in dLGN relay neurons. We will record dLGN relay neurons’ intrinsic properties at various time points following IL-1β eye injections. We will measure the resting membrane potential, input resistance, time constant, average maximum firing rate, spike frequency adaptation, and rheobase to test whether the proinflammatory injection into the vitreous will alter dLGN relay neurons’ intrinsic properties. We hypothesize that injecting IL-1β will lead to modification of the intrinsic properties of dLGN relay neurons. 51 Materials and Methods All animal care and experimental procedures used in this study were approved by the MSU Institutional Animal Care and Use Committee. C57Bl/6 mice of either sex were used. Mice (postnatal day 22-31) were deeply anesthetized using 2-4% isoflurane and kept on a heated pad during the procedure. The sclera and cornea were punctured using a 30 gauge needle and the vitreous was drained to reduce the intraocular pressure. Each eye was injected with 2 μl of IL-1β (10 ng/ml) by using a Hamilton syringe. After the surgery, animals recovered and brain slice recordings were obtained at 24 hours, 48 hours, 72 hours, 1 week, or 2 weeks post injection. Mice were deeply anaesthetized with 2-4% isoflurane, decapitated, and the brain was quickly removed and placed in a cold (<4°C) oxygenated (95 % O2/5 % CO2) slicing solution containing KCl 2.5 mM, NaH2PO4 1.25 mM, MgSO4 10.0 mM, CaCl2 0.5 mM, sucrose 234.0 mM, glucose 10.0 mM and NaHCO3 26.0 mM. Slices (300 μm thick) were cut on a vibrating tissue slicer in the coronal plane. Slices were then incubated in a heated holding chamber (~35°C) containing artificial cerebrospinal fluid solution (ACSF) composed of NaHCO3 26.0 mM, KCl 2.5 mM, glucose 10.0 mM, NaCl 126.0 mM, NaH2PO4 1.25 mM, MgCl2 2.0 mM and CaCl2 2.0 mM for at least 30 minutes and then maintained at room temperature until use. Individual slices were transferred to a submersion recording chamber, perfused with oxygenated ACSF solution at a rate of 3 ml/min and maintained at 32 ± 1°C. Recordings were obtained from dLGN thalamocortical relay neurons. Researchers were blind to the condition (control or IL-1β injections) during recordings. All recordings were 52 performed using the whole cell configuration, and the liquid junction potential (10 mV) was corrected for all measurements. Recording micropipettes had tip resistances of 3-5 MΩ when filled with internal solution containing K-gluconate 117.0 mM, KCl 13.0 mM, MgCl2 1.0 mM, CaCl2 0.07 mM, EGTA 0.1 mM and HEPES 10.0 mM. Data were acquired using a Multiclamp 700B amplifier (Molecular Devices, Union City, CA) in current-clamp mode. Data were filtered at 10 kHz and digitized at 10 kHz using a Digidata 1550 digitizer, and collected using pClamp software (Molecular Devices, Union City, CA). Resting membrane potential (RMP) and input resistance (Rin) were calculated; the RMP was measured befor the hyperpolarization current step injection occurs, when the membrane is at rest (fig 3.1). On the other hand, the Rin was measured from the linear slope of the voltage– current relationship obtained by applying constant current pulses ranging from -200 to +200 pA (1 s duration; fig 3.2). The time constant was measured after injecting a fast (100 ms) and small (-10 pA) current step 20 times, we averaged the data and fit a single order exponential curve to obtain the time constant (fig 3.3). Time constant is the time at which the hyperpolarization current step reaches 63% of the maximum response. An average action potential discharge rate (over 1 s) versus current injection graph was generated for each neuron. To measure spike frequency adaption (SFA), we identified the current step that elicited half the maximum average action discharge rate for each neuron. Using that sweep, we averaged the first three and last three instantaneous action potential frequencies (1/interspike interval) and generated an action potential adaptation ratio (mean last/mean first; fig 3.5). The firing characteristics were measured by using (Clampfit 10.7) analysis software. All data are presented as mean ± 53 standard error of the mean (SEM). All data statistical analysis were done using ANOVA test, where the statistical significance was p<0.05. 54 Results Recordings were obtained from 54 control (non-injected mice) relay neurons, 14 relay neurons after 24 hours following IL-1β injection, 12 neurons after 48 hours, 11 neurons after 72 hours, 8 neurons after 1 week, and 2 neurons after 2 weeks following IL-1β injection. First, we compared the resting membrane potential of dLGN relay neurons between all groups. In the control group, relay neurons had an average resting membrane potential of -69.7 ± 2.5 mV (n=54). While 24 hours following IL-1β injection, relay neurons had an average resting membrane potential of -71.4 ± 2.0 mV (n=14). 48 hours following IL-1β injection, relay neurons had an average resting membrane potential of -70.6 ± 1.0 mV (n=12). 72 hours following IL-1β injection, relay neurons had an average resting membrane potential of -71.9 ± 1.7 mV (n= 11). 1 week following IL-1β injection, relay neurons had an average resting membrane potential of 71.5 ± 1.1 mV (n=8). While 2 weeks following IL-1β injection, relay neurons had an average resting membrane potential of -69.5 ± 2.1 mV (n=2). After running an analysis of variance (ANOVA) with post-hoc Tukey Honestly Significant Difference (HSD) to compare means, the only significant difference was seen between the control and the 72 hours following IL-1β injection (p<0.05, F value: 3.08). This indicates that the resting membrane potential of relay neurons was more hyperpolarized 3 days following IL-1β injection (Fig. 1). 55 Figure 3.1. The relationship between intraocular IL-1β injection and resting membrane potential of dLGN relay neurons. (Top) Population graph of the average resting membrane potentials measured at various time points following IL-1β injection. 72 hours post-injection had a significantly hyperpolarized average resting membrane potential when compared to control mice (Cont.). No other time points tested were significantly different from control values. (Bottom) A representative trace showing where RMP was measured. *p<0.05 56 The average input resistance for the control population was 202.3 ± 114.2 MΩ (n=54), the 24 hours following IL-1β population was 135.3 ± 39.3 MΩ (n=14), the 48 hours following injection population was 162.5 ± 67.6 MΩ (n=12), the 72 hours following injection population was 166.3 ± 91.1 MΩ (n=11), the 1 week following injection population was 163.8 ± 73.6 MΩ (n=8), and the 2 weeks following injection population was 199.3 ± 46 MΩ (n=2). When input resistances were compared between the control population and each post-injection population, they showed no significant differences (ANOVA with post-hoc Tukey HSD, p>0.05, F value: 0.24), indicating that intraocular IL-1β injection does not affect input resistances of dLGN relay neuron within the time periods tested (Fig2). 57 Figure 3.2. The relationship between intraocular IL-1β injection and input resistance of dLGN relay neurons. (Top) Population graph of the average input resistance measured at various time points following IL-1β injection. No significant different was shown between time points tested and control values. (Bottom) A representative trace showing where Rin was measured. 58 We next measured the membrane time constant to test if there are any differences following IL-1β injection. In control neurons, the membrane time constant averaged 27.4 ± 15.4 ms (n= 54), the 24 hours post-injection population averaged 17.2 ± 8.3 ms (n= 14), the 48 hours post-injection populations averaged 18.0 ± 6.0 ms (n= 12), the 72 hours post-injection population averaged 24.1 ± 11.2 ms (n= 11), the 1 week post-injection population averaged 21.1 ± 10.5 ms (n= 8), and the 2 weeks post-injection population averaged 24.0 ± 1.4 ms (n=2). There were no significant changes in the membrane time constant following injection of IL-1β (ANOVA with post-hoc Tukey HSD, p>0.05, F value 0.07), indicating that the time constant of dLGN neurons is unaltered in the time periods tested following intraocular injection of IL-1β (Fig 3). 59 Figure 3.3. The relationship between intraocular IL-1β injection and membrane time constant of dLGN relay neurons. (Top) Population graph of the average time constant measured at various time points following IL-1β injection. No significant alteration was shown between time points tested and control values. (Bottom) A representative trace showing how TC was measured. 60 To further examine neuronal excitability, we determined the average maximum discharge rate. In control neurons, the average maximum discharge rate (over a 1 second current pulse) averaged 75 ± 23 spikes (n=54), the 24 hours population, with an average of 95 ± 49 spikes per second (n=14), the 48 hours populations, with an average of 91 ± 40 spikes per second (n=12), the 72 hours population with an average of 93 ± 40 spikes per second (n=11), the 1 week population with an average of 83 ± 31 spikes per second (n=8), and the 2 weeks population with an average of 174 ± 34 spikes per second (n=2). After analyzing these data using (ANOVA with post-hoc Tukey HSD), the 2 weeks population showed a significant increase in average maximum discharge rate compared with all other populations (p<0.05, F value 0.0016), however it is important to note only two samples were tested (Fig. 4). 61 Figure 3.4. The relationship between intraocular IL-1β injection and the average action potential discharge rate of dLGN relay neurons per second. (Top) Population graph of the average action potential discharge rate measured at various time points following IL-1β injection. No significant alterations were shown between time points tested and control values. (Bottom) A representative trace showing how the average action potential discharge rate was measured. 62 In order to quantify the magnitude of action potential frequency adaptation, we measured the adaptation ratio (see methods) and compared them across populations. The control population ratio averaged 0.34 ± 0.22 (n=54), 24 hours post IL-1β injection ratio averaged 0.21 ± 0.09 (n=14), 48 hours post injection group had an averaged ratio of 0.28 ± 0.17 (n=12), 72 hours post injection group had an averaged ratio of 0.25 ± 0.11 (n=11), the 1 week post injection group had an averaged ratio of 0.39 ± 0.21 (n=8) and the 2 week post injection group had an averaged ratio of 0.26 ± 0.06 (n=2). No significant changes were seen between the control and post injection time groups (ANOVA with post-hoc Tukey HSD, p>0.05, F value: 0.15), indicating that spike frequency adaptation is unaltered in the time points tested following IL-1β injections (Fig. 5). 63 Figure 3.5. The relationship between intraocular IL-1β injection and spike frequency adaptation (SFA) of dLGN relay neurons. (Top) Population graph of the average SFA measured at various time points following IL-1β injection. No significant alteration was shown between time points tested and control values. (Bottom) A representative trace showing how SFA was measured. 64 We next measured the rheobase to test whether the inflammatory response is changing the amount of current necessary to produce action potential discharge in the relay neurons. The average current that elicited action potential discharge for the control population was 54 ± 34 pA (n=54). The average rheobase for 24 hours post-injection population was 55 ± 25 pA (n=14), the average rheobase for 48 hours post-injection population was 67 ± 52 pA (n=12), the average rheobase for 72 hours post-injection population was 57 ± 34 pA (n=11), the average rheobase for 1 week post-injection population was 56 ± 19 pA (n=8) and the average rheobase for 2 weeks post-injection population was 90 ± 42 pA (n=2). No significant changes were noted (ANOVA with post-hoc Tukey HSD, p>0.05, F value: 0.67), indicating that IL-1β injection does not alter the rheobase in the time periods tested. (Fig. 6). 65 Figure 3.6. The relationship between intraocular IL-1β injection and rheobase of dLGN relay neurons. (Top) Population graph of the average rheobase measured at various time points following IL-1β injection. No significant alteration was shown between time points tested and control values. (Bottom) A representative trace showing how rheobase was measured. 66 Discussion In this study we have shown that intraocular injections of IL-1β alter intrinsic properties of dLGN relay neurons. These alterations in the intrinsic properties of relay neurons occur at 72 hours and 2 weeks were the 2 weeks have only 2 cells, after IL-1β injection. These findings indicate that after the injections of IL-1β, the inflammatory response might lead to changes in the RGCs, three days after injection, it hyperpolarized the resting membrane potential of dLGN neurons. It also indicates that there might be a chronic inflammation affecting average maximum firing frequency after 2 weeks post injection, but with the number of cells analyzed it not enough to make a definite conclusion, this alteration might be related to the hyperexcitability due to either the excitotoxicity of RGC or the RGC apoptosis process. The Hyperexcitability findings obtained two weeks after IL-β injection might suggest that glutamate released during RGCs apoptosis process, could lead to the hyperexcitability of the dLGN neurons. This neuronal hyperexcitability might be the key for visual problems associated with diabetic retinopathy. These findings could direct us in many directions; first we need to collect more data two weeks post injection, to confirm our outcome. Then we could lengthen the time between injections and recordings following IL-1β, to test whether the excitability seen in this study will continue to increase and at what point does it return to control levels. We also could use higher dosages, which would accelerate RGC cell death and in turn increase the influence on dLGN. We also can test whether the changes in intrinsic properties are dose dependent or not. Since it is known that diabetic retinopathy is a chronic condition, we can add more than one set of 67 injections, meaning that we give 3 to 4 injections with a set time between injections to extend the inflammatory response period, and test if there will be any differences. We also could do the same experiment and study IL-1β injection effect on dLGN interneurons, since they are deeply involved in the regulation of retinal output to the dLGN, and test whether it will alter its intrinsic property. From this experiment, we conclude that intraocular injection of pro-inflammatory cytokine IL-1β, can alter dLGN neurons intrinsic properties. Studies have shown that retinal cell death, which could alert RGC output, could be prevented with the combination of many therapeutic strategies, such as the use of anti-inflammatory agents, which could significantly decrease the apoptosis rate, another way to prevent retinal cell death is by inhibiting glutamate, by using NMDA receptor antagonist, and one of the useful methods is the use of cannabidiol as neuronal protective therapy to prevent retinal cell death (Kern and Barber, 2008). 68 REFERENCES 69 REFERENCES Adamis AP and Adrienne BJ. Immunological mechanisms in the pathogenesis of diabetic retinopathy. Seminars in Immunopathology. 2008: 30, 65-84. Antonetti DA, Barber AJ, Bronson SK, Freeman WM, Gardner TW, Jefferson LS, Kester M, Kimball SR, Krady JK, LaNoue KF, Norbury CC, Quinn PG, Sandirasegarane L and Simpson IA. Diabetic Retinopathy. Seeing beyond glucose-induced microvascular disease. Diabetes. 2006: 55, 2401-2411. Barber AJ, Lieth E, Khin SA, Antonetti DA, Buchanan AG and Gardner TW. Neural apoptosis in the retina during experimental and human diabetes. The Journal of Clinical Investigation. 1998: 102, 783-791. Ciulla TA, Amador AG, and Zinman B. Diabetic retinopathy and diabetic macular edema. Diabetes Care. 2003: 26, 2653-2664. Eskan MA , Benakanakere MR, Rose BG, Zhang P , Zhao J, Stathopoulou P, Fujioka D and Kinane DF. Interleukin-1β modulates pro-inflammatory cytokine production in human epithelial cells. Infection and Immunity. 2008: 76, 2080-2089. Gardner TW, Antonetti DA, Barber AJ, LaNoue KF and Levison SW. Diabetic retinopathy: more than meets the eye. Survey of Ophthalmology. 2002: 47, 253-262. Kastner S, Schneider KA, and Wunderlich K. Beyond a relay nucleus: neuroimaging views on the human LGN. Progress in Brain Research. 2006: 155, 125-143. Kempen JH , O'Colmain BJ , Leske MC , Haffner SM , Klein R , Moss SE , Taylor HR and Hamman RF. The prevalence of diabetic retinopathy among adults in the United States. Archives of Ophthalmology. 2004: 122, 552-563. Kern TS and Barber AJ. Retinal ganglion cells in diabetes. The Journal of Physiology. 2008: 568, 4401-4408. Kohner EM, Patel V, and Rassam SM. Role of blood flow and impaired autoregulation in the pathogenesis of diabetic retinopathy. Diabetes. 1995: 44, 603-607. Michelle H and Akpek EK. Dry Eye: An inflammatory ocular disease. Journal of Ophthalmic & vision research. 2014: 9, 240–250. 70 Lieth E, Gardner TW, Barber AJ and Antonetti DA. Retinal neurodegeneration: early pathology in diabetes. Clinical and Experimental Ophthalmology. 2000: 28, 3–8. Penn JS, Madan A, Caldwell RB, Bartoli M, Caldwell RW, and Hartnett ME. Vascular endothelial growth factor in eye disease. Progress in Retinal and Eye Research. 2008: 27, 331-371. Peron SP and Gabbiani F. Role of spike-frequency adaptation in shaping neuronal response to dynamic stimuli. Biological Cybernetics. 2009: 100, 505-520. Sheetz MJ and King GL. Molecular understanding of hyperglycemia’s adverse effect for diabetic complications. The Journal of the American Medical Association. 2002: 288, 2579-2588. Tang J and Kern TS. Inflammation in diabetic retinopathy. Progress in Retinal and Eye Research. 2011: 30, 343–358. Valenciano AL, Boya P and Rosa EJ. Early neural cell death: numbers and cues from the developing neuroretina. The International Journal of Developmental Biology. 2009: 53, 15151528. Williams R, Airey M, Baxter H, Forrester J, Kennedy-Martin T and Girach A. Epidemiology of diabetic retinopathy and macular oedema: a systematic review. Eye. 2004: 18:963-983. Wong RL. Retinal waves and visual system development. Annual Review of Neuroscience. 1999: 22, 29-47. 71 Chapter 4 Summary and future direction Summary The brain is the organ that controls all other body functions, it controls every single movement and every single thought we do, and its regions are well connected. Both FXS and diabetic retinopathy are pathological conditions that have significant influences in the brain. In diabetic retinopathy, defect in the retinal RGCs could have an effect on the thalamus relay neurons and it even might have a downstream effect on the cortical pyramidal neurons. On the other hand, individuals with FXS, which is manifested in the brain, are experiencing many neurological and behavioral defects. Up to date there is no definite treatment to prevent either condition. The lack of cure as well as the expansion in their influences gave both FXS and diabetic retinopathy the attention of scientist to fully understand those conditions and to participate to find their cures. FXS is associated with many cognitive, physical and biological defects. It occurs in a ratio of 1:4000 in males and 1:8000 in females making it the most common form of mental retardation. A genetic correction of the hypermethylation in the Fmr1 gene could be possible someday, leading to a normal production of FMRP. Also, treating the mGluRs influence on the disease might be a solution in the near future, by finding the right pharmacological agent that can prevents the overactivation of group I mGluR, leading to the reduction of excitation associated with FXS, without affecting the numerous functions of mGluR. Finding such treatments might lead to the regulation of synaptic protein synthesis, which will solve many 72 hyperexcitability issues associated with FXS. Finding a treatment for FXS is the dream of many scientists, who spent their lives trying to make a miracle. On the other hand, with our modern lifestyle, and with the fast and uncontrolled rise in diabetic patients, the occurrence of diabetic associated blindness will eventually rise. This increase will have an economic as well as productivity influences. When blindness increases, affected people may require more care and may hospitalization. Diabetic retinopathy could be controlled when detected in the early stage, to slow its prognosis, but not to revers it, and it is untreatable, up to date, when it gets to its advanced stage. Continues vision monitoring is critical for diabetic patients, and any visual changes could be a sign for developing diabetic retinopathy. The main reason for diabetic retinopathy development is its association in the RGCs death, preventing the death of RGCs could significantly limit the defects in diabetic retinopathy. RGCs death is the result of glia and microglia cells activation, which leads to the activation of caspase 1 as well as production of many inflammatory cytokines and the production of glutamate. The excess glutamate production as well as the activation of caspase 1 will lead to the RGCs death. With more research and work in both FXS and diabetic retinopathy, we will eventually be able to treat them and prevent their influence on humanity. 73 Future directions after the postsynaptic actions of mGluR in Fmr1 KO mice study Since the activation of mGluRs is inducing alterations in plasticity in different brain regions, and after our results that showed no alterations in short-term induction of both group I and II mGluR agonist, we could try the induction of a general mGluR agonist for longer than (60s) and test its effect on Fmr1 KO mice and compare it to WT. This might show that the binding affinity of the mGluR is altered in the Fmr1 KO mice after a set time, and this alteration could be shown by increasing the drug application time. We also could record from interneurons instead of excitatory pyramidal; this alteration in excitability could be driven by an exaggerated inhibition. We might find that the Fmr1 KO mice are having significant reduction in the inhibitory circuit, leading to the hyperexcitability manifestation seen in FXS. On the other hand, we could record from excitatory neurons in deep layers visual cortex, where we previously showed the alteration in plasticity, to support either outcome. We also could record from inhibitory interneurons in visual cortex, to test wither the hyperexcitability is due to reduced inhibition. The conformation of the defects in mGluR would be significantly helpful for the full understanding of FXS and it could lead to the production of suitable pharmacological agents that could preserve the crucial functions of mGluR in FXS. With the huge amount of work and publications in FXS, and with the full understanding of the condition will make the prevention of FXS become possible. 74 Future direction after the IL-1β intraocular injection and its alteration on the dorsal lateral geniculate nucleus relay neurons Based on the potential changes two weeks after intraocular injection of IL-1β, it is critical to lengthen the time between injection and recording in order to confirm the alteration seen in the average maximum firing discharge as well as increase the number of data in the two weeks post injection group. It is also beneficial to start recording from the RGCs and determine the impact of intraocular injections of IL-1β on the intrinsic activity of RGCs. When recording directly from the RGCs, we could determine whether they are altered, and if they are, what the time point of RGC alteration, which can influence neurons of the dLGN. Currently we are testing the influence of an acute inflammatory condition, after the outcome of this study, it may be more affective to use multiple injections (3-4 instead of one) to produce a chronic condition rather than an acute one, since diabetic retinopathy is a chronic disease. We then could use the 3 or 4 week post injection mice to test the influence of the chronic inflammatory response on the RGCs as well as the neurons of the dLGN. Using higher IL-β concentration is going to be helpful to determine the influence of IL-1β on the RGC health. By injecting higher concentration, we are accelerating the RGC cell death, which could allow us to clearly identify the downstream effect of that on the dLGN relay neurons. In this study, we have collected all eyes after mice decapitation and they were kept in … for farther assessment, it is significant to look at these eyes and assess the degree of damage to 75 the RGCs in each post-injection population, if there are any damages, which would be beneficial to identify the affect associated with the dosage we used and gives us a window for IL-1β concentration manipulation. 76