Biochimica et Biophysica Acta 1812 (2011) 333–334 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Editorial Preface: Zebrafish Models of Neurology☆ Forming hypotheses about the Molecular Basis of Neurological Disease has traditionally been based on human genetics, biochemistry and molecular biology, whereas testing these hypotheses requires us to turn to cell culture and animal models. Zebrafish have emerged as a tractable platform to complement and bridge these paradigms in many diseases, allowing integration of human genetics and cell biology within the in vivo setting. Perhaps it is in Neurology and Neurodegeneration, however, where zebrafish have special promise to make contributions. Four special advantages of zebrafish to Neurology are worth highlighting here: efficient in vivo tests for genetic interactions, conserved cellular architecture of the CNS, the ability to create genetically mosaic animals, and tractable behaviour and electrophysiology to assess experimental interventions. Linkages between Neurodegenerative diseases that were once considered distinct are being recognized and celebrated as a path forward to understanding the molecular basis of disease spread through the CNS. One example includes interactions between notorious proteins in Alzheimer Disease (AD) and Prion Disease, such that the Prion Protein may now be viewed as a receptor for Aβ oligomers. The abundance of such interactions between Neurological Diseases is growing, as protein interactome studies identify many players in common, compelling the field to integrate an increasing list of genetic interactions into their models. The ability to efficiently knockdown and overexpress gene products in vivo is propelling zebrafish to prominence at a rate commensurate with the complexity of these genetic interactions. On the other hand, various researchers have recently noted intriguing parallels between prion diseases and the protein misfolding in other neurological diseases. In particular, questions are being raised as to the ability of misfolded proteins (including Tau and Aβ) to induce the misfolding of normally folded protein, leading to a prion-like domino effect and accumulation of protein dysfunction/aggregation that are hallmark disease pathologies. Consistent with such findings is the apparent spreading of disease pathology from a nucleation centre through the CNS during clinical disease progression. Other celebrated potential prion-like proteins include α-synuclein, polyglutamine proteins and SOD1 misfolding in ALS. Attention toward this prionoid nature of various neurological diseases is substantial, in large part because it identifies a novel therapeutic target: the mechanism of disease spreading. Unfortunately, the field knows remarkably little about disease spreading, with concepts in high-profile reviews centering on Tunneling Nanotubes and exosome or synaptic release. Perhaps this fundamental lack of knowledge regarding disease spread through the CNS is the consequence of the tools available to neurologists and neuroscience researchers. The mechanisms underlying AD, TSEs, and ALS have ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. 0925-4439/$ – see front matter © 2010 Published by Elsevier B.V. doi:10.1016/j.bbadis.2010.11.004 benefitted primarily from biochemistry/ biophysics of protein aggregation, cell culture work and mouse models; These paradigms are challenging to implement in a fashion that tells us how disease in one portion of the CNS spreads to another. Zebrafish are positioned brilliantly to fill this gap, not only because of its efficiency as a genetic and neuroscience model, but because zebrafish have a long history as a platform for assessing non-cell-autonomous effects of protein perturbations in the in vivo setting. Thus it is common for zebrafish biologists to ask if changing gene abundance or character in one group of neurons is able to affect the fate and function of neighboring cells. Creative and ambitious deployment of this ability will complement established tools to promise a deep understanding of how disease spreads between cells. A tantalizing side-benefit is the ability to deploy the fish in high-throughput screens of protein and small molecule effectors that represent therapeutic targets. Finally, it is important to underline the cellular context of disease spreading during Neurolodegeneration. As disease spreads from cell to cell, the molecular basis of this spread is fundamentally influenced by the cellular relationships. Importantly, The zebrafish has an architecture of CNS cells that is conserved with that of humans. Vasculature, blood-brain-barriers, and ventricles are all present from a young age, permitting tissue-level spread of molecules fundamental to disease. Perhaps more importantly, the cells of the CNS develop in their normal environment, becoming polarized, interconnected and ensheathed in radial glia to form ion sinks. The cover image selected is meant to highlight that this organization exists even in young embryos. Certainly our understanding of hallmark synaptic dysfunction during Neurodegeneration must develop from paradigms where synapses are formed and maintained with their typical glial partners intact. So although zebrafish is not yet as advanced in its toolkit as Drosophila, nor as tractable as cell culture, it has a relevant cellular architecture and genetic efficiency that make the paradigm a very attractive in vivo complement to studies in mouse models. This issue of BBA: Molecular Basis of Disease has invited reviews from leading experts to consider the utility of zebrafish to Neurology. These experts span from Neuroscientists who study fundamentals of zebrafish Biology, through to clinical Neurologists who recognize the zebrafish as a relevant and potent paradigm to bring their hypotheses to bedside in as timely a manner as possible. The first contributions have sought to critically analyze previous work regarding particular Neurodegenerative diseases and how zebrafish have contributed to this understanding. The subsequent contributions consider technical opportunities and challenges of using zebrafish genetics and behavioural outputs in Neurology, and the issue ends with the promise provided by zebrafish to understand regeneration of the CNS for repair. I am grateful to each of these contributors, reviewers, and the BBA staff for their hardwork in communicating these important and stimulating ideas. 334 Editorial Allison is a neuroscientist with a special interest in understanding processes of neurodegeneration and neural regeneration. Towards this goal, Dr. Allison's research integrates molecular biology, biochemistry, proteomics, and electrophysiology to further develop zebrafish as a potent animal model. Particular emphasis of his research has been developing fish as effective models of neurodegenerative disease. He completed a Doctoral Program in 2004 with Professor Craig Hawryshyn at the University of Victoria, Canada, where he developed a unique model of neuronal degeneration and regeneration. This work was funded by fellowships from the Alzheimer Society of Canada and the Canadian Institute of Health Research Institute of Aging. Dr. Allison went on to complete an NSERC Post-Doctoral fellowship at the University of Michigan in the laboratory of Collegiate Professor Pamela Raymond. Dr. Allison's work at the University of Michigan created transgenic and mutant zebrafish has revealed genes important in the generation, positioning, and regeneration of neurons. Dr. Allison joined the Centre for Prions and Protein Folding Diseases as an Assistant Professor at the University of Alberta, Canada in 2008. His appointment is through the Department of Biological Sciences in the Faculty of Science, with a cross-appointment in the Department of Medical Genetics in the Faculty of Medicine and Dentistry. Dr. Allison is contributing to expanding the largest zebrafish facility in Western Canada and also maintains active research in the field of retina development and regeneration. His current research efforts focus on creating transgenic and mutant zebrafish to address hypotheses of prion protein's normal function and to dissect the role of various prion protein domains. Other efforts use similar methods to examine the amyloid hypothesis of Alzheimer Disease. His Recent work is funded by a Recruitment Grant from PrioNet Canada and the Alberta Prion Research Institute and by a Young Investigators Award from the Alzheimer Society of Canada. W. Ted Allison University of Alberta, Centre for Prions and Protein Folding Diseases, Departments of Biological Sciences & Medical Genetics, Edmonton, Alberta, Canada T6G 2E9 E-mail address: [email protected]. Tel.: +1 780 492 4430; fax: +1 780 492 9234. 10 November 2010 Available online 27 November 2010 Biochimica et Biophysica Acta 1812 (2011) 335–345 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Zebrafish models for the functional genomics of neurogenetic disorders☆ Edor Kabashi a,b, Edna Brustein b, Nathalie Champagne a,b, Pierre Drapeau b,⁎ a b Center for Excellence in Neuromics, CHUM Research Center and the Department of Medicine, Université de Montréal, Montréal, QC, Canada Department of Pathology and Cell Biology and Groupe de recherche sur le système nerveux central, Université de Montréal, Montréal, QC, Canada a r t i c l e i n f o Article history: Received 4 December 2009 Accepted 22 September 2010 Available online 29 September 2010 Keywords: Zebrafish Human diseases Neurogenetic diseases Neurodegenerative disorders Neurodevelopmental diseases Animal models Genomics Human genetics Trangenic animals a b s t r a c t In this review, we consider recent work using zebrafish to validate and study the functional consequences of mutations of human genes implicated in a broad range of degenerative and developmental disorders of the brain and spinal cord. Also we present technical considerations for those wishing to study their own genes of interest by taking advantage of this easily manipulated and clinically relevant model organism. Zebrafish permit mutational analyses of genetic function (gain or loss of function) and the rapid validation of human variants as pathological mutations. In particular, neural degeneration can be characterized at genetic, cellular, functional, and behavioral levels. Zebrafish have been used to knock down or express mutations in zebrafish homologs of human genes and to directly express human genes bearing mutations related to neurodegenerative disorders such as spinal muscular atrophy, ataxia, hereditary spastic paraplegia, amyotrophic lateral sclerosis (ALS), epilepsy, Huntington's disease, Parkinson's disease, fronto-temporal dementia, and Alzheimer's disease. More recently, we have been using zebrafish to validate mutations of synaptic genes discovered by large-scale genomic approaches in developmental disorders such as autism, schizophrenia, and non-syndromic mental retardation. Advances in zebrafish genetics such as multigenic analyses and chemical genetics now offer a unique potential for disease research. Thus, zebrafish hold much promise for advancing the functional genomics of human diseases, the understanding of the genetics and cell biology of degenerative and developmental disorders, and the discovery of therapeutics. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Introduction Zebrafish are used as a model for a wide variety of human diseases, including cancer, cardiovascular disorders, angiogenesis, hemophilia, osteoporosis, diseases of muscle, kidney and liver, and, last but not the least, disorders of the central nervous system. In recent years, zebrafish have been used to study neurodegenerative disorders. Here we review zebrafish models of CNS diseases with an emphasis on studies of degenerative neurological diseases. We also discuss our recent efforts in extending these approaches to psychiatric disorders of development. Although unlikely to yield accurate models for complex psychiatric disorders, genetic insights from studies of zebrafish are pertinent in helping to identify relevant molecular and cellular mechanisms of human pathology and, at this level, promise insight to the development of therapeutics. In particular, the in vivo biological validation of variants identified in human genetic and genomic studies provides an important step in defining the pathological nature of mutations. ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Corresponding author. Department of Pathology and Cell Biology, Université de Montréal, C.P. 6128, Succ. Centre-ville, Montréal, Québec H3C 3J7, Canada. Tel.: +1 514 343 7087; fax: +1 514 343 5755. E-mail address: [email protected] (P. Drapeau). 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.09.011 Apart from being a vertebrate with common organs and tissues such as a brain and spinal cord with conserved organization, the attractiveness of zebrafish as a model lies in its biology and genetics. Zebrafish have large clutches of externally fertilized and transparent eggs, which develop rapidly and in synchrony, with neurogenesis starting around 10 h post-fertilization (hpf), synaptogenesis and the first behaviors around 18 hpf and hatching around 52 hpf. Within 1 day of development, many pertinent features of the CNS appear and can be studied in relatively simple populations of identifiable neurons. For example, the spinal cord is divided into 30 somites each containing fewer than a dozen cell types that form relatively simple circuits. The rapid development of the zebrafish embryo allows for the study of embryonic-lethal mutations as larvae can survive on their yolk for up to a week, allowing studies of gene expression and function throughout early developmental stages. In addition to its advantageous biology, the second major advantage of the zebrafish as a model is the simplicity and effectiveness of manipulating gene expression for cell biological observations in living embryos with relevance to human pathology. The zebrafish genome is sequenced and, although not completely annotated, over 80% of gene structures are available and show a high degree of synteny (nearest neighboring genes on the chromosomes) across vertebrate species as well as 50–80% homology with most human sequences. Homologs for most human genes can be identified in silico and, as described below, can be manipulated by gain- and loss- 336 E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 of-function approaches. However, the identification of a human homolog in zebrafish is complicated by the fact that a large number of genes are duplicated. This phenomenon is explained by whole genome duplication that occurred during evolution [1]. The amino acid identity compared to the human gene is always higher for one of the two genes duplicated, indicating that the duplication appeared before the species divergence. Therefore, when a gene is knocked down, it is important to consider that the duplicated gene can compensate for the loss of function of the targeted homolog. Also, the function could be partitioned between the two genes or simply lost for one of the genes. Comparative sequence analysis allows identifying the zebrafish homolog of a human gene. Three major genome browsers, NCBI (www.ncbi.nlm.gov), Ensembl (www.ensembl.org) and UCSC (/genome.ucsc.edu/), can be use to search homologous gene sequences. In many cases, the homologous gene predictions can be found directly from either HomolGene (NCBI) or Orthologue Prediction (Ensembl). We have found that about half of the hundred or so human synaptic genes on chromosome X (described later) have a predicted fish ortholog, compared to ~ 30% in Drosophila and Caenorhabditis elegans using the same databases. If the homologous gene is not annotated or if the transcript prediction is not complete, a BLAST search using nucleotide or protein sequence against the whole genome assemblies (especially scaffolds based on BAC sequencing) can allow the identification of homologous genes. TargetP and SignalP algorithms (www.cbs.dtu.dk/services/) can be used for signal peptide and cleavage site prediction. An additional phylogenetic analysis and multiple alignments of protein sequences (Clustal W) help to confirm the orthology. The syntenic regions can also be compared using Multi Contig View (Ensembl). For additional information, the Sanger Institute (www.sanger.ac.uk) has excellent resources about zebrafish sequence analysis. The sequence identity between zebrafish genes and their human homologs is higher for ubiquitously expressed genes than for highly specific genes such as some synaptic genes. The divergence is mainly found in domains important for membrane targeting as signal peptides and often in intracellular domains that contain signaling motifs. Nonetheless, in general, the exon/intron boundaries and regulatory sequences as well as many other types of functional domains are evolutionarily conserved between the species. As a simple starting point, the zebrafish homologs can be targeted for knockdown by injection of selective anti-sense morpholino oligonucleotides (AMOs) [2]. ZFIN (zfin.org) has the principal database about AMO gene knockdown and phenotype. The parameters to consider in AMO design are the same as that for any oligonucleotide molecule: CG content, stem and loop structure, and the oligo length [3]. Gene-Tools Inc. (see www.gene-tools.com) offers an AMO design service to the researchers purchasing their morpholinos. We recommend always checking if the AMO can bind to off-target sequences by doing a BLAST against NCBI or Ensembl database. AMOs can be designed to target the ATG start site to block translation initiation or they can be used to block splice sites to produce truncated transcripts. The latter have the advantage of sparing maternal transcripts and permit RT-PCR amplification and quantification of the gene knockdown, which is particularly useful when an antibody is unavailable. Further, the splice AMO can be used to mimic splicing defects, nonsense or truncation mutations related to disease. AMOs resist degradation by ribonucleases and are stable for several days, and in our hands when adequately controlled, they usually generate selective phenotypes (discussed below). However, AMOs must be injected intracellularly as they are more of a pharmacological than a genetic tool, although they are more effective in gene knockdown in zebrafish than interfering RNAs. Intracellular injection is feasible in the zebrafish embryo during the first 1–2 hpf as the blastomeres are open to the yolk prior to the 4th cell division. Freehand injection into the yolk near the blastomere border (Fig. 1) allows for the injection of a hundred eggs or so in 1–2 h, yielding a sufficient number of treated embryos to permit statistical analyses of phenotypes and thus providing a genetic read-out often only 1 day later. Replacing targeted zebrafish genes with human genes can for many genes be done as simply as by injection of human mRNA simultaneously with the AMO as the latter is designed to specifically target the zebrafish and not the human sequence. We suggest using the zebrafish homologous cDNA if your gene of interest has a highly localized expression pattern like that of synaptic genes. Otherwise, human, mouse, or rat cDNA can be used and easily obtained commercially. NCBI gives the link to cDNA clones for certain genes. The gene can be subcloned in expression vectors like pCS2 [4] or pcGlobin [5] to synthesize the 5′-capped mRNA that will be injected in embryos. The pcGlobin vector has the advantage to have 5′UTR and 3′ UTR from Xenopus, and we also note that pcGlobin gives a better yield of mRNA. The cDNA can be tagged to MYC, flag, or HA epitopes to follow the expression in fish. However, large DNA constructs such as plasmids often yield chimeric expression patterns in limited numbers of cells presumably as not all dividing progenitors retain the large constructs. About half of the human genes we have tested permit partial rescue of knockdown phenotypes, which then allows for comparison between wild-type human mRNA and mRNA bearing disease-related mutations. Thus, in the space of a few weeks, each human variant can be validated in knockdown embryos once a clear phenotype is recognized. Finally and as described below, recent transgenic technologies permit the creation of stable, targeted (cellspecific), and inducible lines for better spatially and temporally controlled and more detailed analyses of transgenic function. 2. Degenerative brain disorders Degenerative disorders of the brain, including Alzheimer's, Parkinson's, and Huntington's diseases, are the first neurogenetic diseases to have been studied using zebrafish [6,7]. Amyloid-beta plaques are the strongest biological marker of Alzheimer's disease and the production of amyloid-beta is regulated by the presenilins 1 and 2 (PS1, PS2). Surprisingly, wt zebrafish zPS1 was found to promote aberrant amyloid-beta42 secretion when expressed in HEK 293 cells, like the abnormal secretion associated with human PS1 mutations, and mutation of one residue in zPS1 abolished this activity [8]. Truncation of zPS1 due to loss of exons 8 and 9 upon injection of splice acceptor AMOs had a dominant-negative effect by increasing PS1 expression [9] and possibly the expression of other Alzheimer's-related genes [10]. Zebrafish zPS2 is expressed at later stages and is regulated by zPS1 [11,12], but its role in amyloidogenesis in zebrafish is unclear. Zebrafish possess two homologues of amyloid precursor protein (APP) with 63–66% homology to their human counterpart [13], and their double knockdown causes a convergence–extension developmental defect that is rescued by human APP mRNA but not by a Swedish mutation related to Alzheimer's disease [14]. The zebrafish APPs possess a functional gamma-secretase complex to produce amyloid-beta [15] and inhibition of gamma-secretase [16] or knockdown of Pen-2 [17], another member of the presinilin complex, blocks Notch signaling to produce a severe neurogenic phenotype. Early studies also examined the tau protein in zebrafish as tau is present in human neurofibrillary tangles, mutations of tau are implicated in dementia and tau may also contribute to Alzheimer's disease. The zebrafish studies have shown that expression of human tau (driven transiently upon injection of a construct with a neural-specific variant of the GATA2 promoter) results by 2 days in disruption of cytoskeletal structure, tau trafficking, and hyperphosphorylated fibrillar tau staining [18]. More recently, stable transgenic zebrafish expressing mutated P301L human tau and a fluorescent reporter (driven panneuronally from the HuC promoter) have permitted in vivo imaging of defective axonal growth, tau hyperphosphorylation, and the screening of novel therapeutic molecules [19]. These recent developments with zebrafish tau and amyloid open the possibility of further screens for novel therapeutics that, e.g., decrease the hyperphosphorylation of tau E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 337 Fig. 1. Rescue of knockdown phenotypes by human mRNA allows for the validation of disease-related variants. in HuC transgenics [19] or amyloid-beta load in amyloid transgenics [20,21]. Research into Parkinson's disease has also used zebrafish as a model [22]. Zebrafish embryos treated with MPTP, a neurodegenerative chemical that reproduces some of the effects of idiopathic Parkinson's disease in mammalian models, demonstrated a loss of dopaminergic neurons, which could be rescued using the monoamine oxidase-B inhibitor deprenyl [23–25]. Half a dozen genes have been identified in Parkinson's disease [26], and several of these have been studied to date in zebrafish. For example, in zebrafish, the mRNA for the ubiquitin processing gene UCH-L1 was detected by 1 day in the ventral region of the midbrain and hindbrain and in the ventral diencephalon where it was co-expressed with markers of dopaminergic neurons [27]. The zebrafish DJ-1 protein has high (N80%) homology with human and mouse DJ-1 (of unknown function) and is expressed throughout the body [28]. Knockdown of DJ-1 in the zebrafish did not affect the number of dopaminergic neurons, but zebrafish embryos were more susceptible to oxidative stress and had elevated SOD1 levels, while simultaneous knockdown of DJ-1 and p53 caused dopaminergic neuronal loss. Recently zebrafish Parkin, another ubiquitin processing gene, was identified (62% homology with human PARKIN) and AMO abrogation of its activity leads to a significant decrease in the number of ascending dopaminergic neurons in the posterior tuberculum, which is homologous to the substantia nigra in humans [29]. Screens for compounds promoting (like MPTP) or preventing (such as caffeine) the Parkinsonian phenotype in zebrafish [30] should now be facilitated by the availability of an enhancer trap line expressing GFP from the vesicular monoamine transporter 2 promoter [31]. In zebrafish, the huntingtin (Htt) gene implicated in Huntington's disease has a predicted 70% identity with the human protein [32] and its knockdown disrupts a number of features [33] and causes massive neuronal apoptosis due to reduced BDNF expression by 1 day of development but not earlier [34]. It was recently observed that injecting mRNA coding for the N-terminal fragment of Htt with different length polyQ repeats led to developmental abnormalities and apoptosis in the embryos as early as 1 day later [35]. Embryos expressing a Q102 mRNA [35] or a Q56 plasmid [36] developed inclusions in the cytoplasm. These Huntington's disease models are being used to screen for novel therapeutics that could prevent aggregate formation and clearance, such as anti-prion compounds [35], or embryonic death, such as blockers of autophagy [37]. Another CNS disorder that has been studied recently using zebrafish is epilepsy [38] as clonus-like seizures are induced by convulsant agents [39–42], which has permitted forward mutagenesis screens and the isolation of seizure-resistant mutants [43]. Zebrafish are now being used to screen for compounds that suppress seizures [44]. Furthermore, injection of mRNA with a mutation of the human PRICKLE1 gene implicated in epilepsy disrupts normal function when overexpressed in zebrafish [45], demonstrating the usefulness of zebrafish for validating mutations of human genes causing brain disorders. Our group has recently collaborated on the discovery of the MEDNIK syndrome, a rare and severe autosomal recessive neurocutaneous disorder manifested by mental retardation, enteropathy, deafness, neuropathy, ichthyosis, and keratodermia that is often lethal [46,47]. Affected individuals bear an A to G mutation in acceptor splice site of exon 3 of the AP1S1 gene, which leads to a premature stop codon (Fig. 2A [47]). The AP1S1 gene encodes the small subunit σ1A of the first (AP-1) of four ubiquitous clathrin adaptor proteins [48–51]. Each one of the adaptor protein complexes is assembled from four subunits, and the σ subunit, the one affected in the MEDNIK syndrome, is part of the AP complex core and is suggested to contribute to its stabilization. Also, together with the μ subunit, the σ subunit is possibly involved in protein cargo selection [50,51]. To demonstrate that the mutation in the human AP1S1 gene indeed alters the biological function of this gene and underlies the MEDNIK syndrome, we knocked down the function of the homologous Ap1s1 gene in zebrafish. The zebrafish Ap1s1 protein shares 91% identity with the human protein. To inhibit zebrafish Ap1s1 mRNA translation, two types of AMOs were designed (Fig. 2A [47]). The first AMO targeted the N-terminal, while the second AMO targeted the acceptor splice site of intron 2 (Fig. 2A), imitating the mutation found in MEDNIK patients. Both AMOs caused similar, severe morphological and behavioral deficits. The 48 hpf KD larva was smaller and had reduced pigmentation compared to the wild-type (WT) larva. Further, blocking Ap1s1 translation caused disorganized skin formation in general, particularly affecting fin morphology (Fig. 2B; for further information on the epithelial disorders, see Ref. [47]). In addition to the morphological deficits, the 48 hpf KD larva responded abnormally 338 E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 to touch by coiling the tail instead of swimming away. The compromised touch response of the KD larva could be attributed to a massive reduction (by half) in the spinal neuron population, specifically due to a loss in the interneuron population. The specificity of the AMO effect was further confirmed both by Western blotting and immunolabeling analysis (Fig. 2C and D). Most importantly, we could rescue the knockdown phenotype by co-injecting the AMO with human AP1S1 wild-type mRNA (Fig. 2B and D), which is not targeted by the Ap1s1 AMO. In contrast, co-injection of AMO with mutated human AP1S1 mRNA missing exon 3 (Fig. 2A) failed to rescue the skin and behavioral deficits, suggesting loss of function of this truncated form of protein (Fig. 2C) and confirming the pathogenic nature of the mutation found in MEDNIK patients. Further, co-injection of AMO and an additional human AP1S1 mRNA isoform found in MEDNIK patients, containing a cryptic splice acceptor site located 9 bp downstream of the start of the 3rd exon, rescued the phenotype (Fig. 2A). The latter result suggests that the predicted in frame protein, lacking only 3 amino acids (Fig. 2C), is functional and may explain the viability of the MEDNIK patients. However, the low expression level of the alternatively spliced RNA (~10%) is insufficient to sustain normal development and function, further emphasizing the importance of AP1S1 in normal development. Our in vivo zebrafish study of Ap1s1 function revealed its importance for appropriate neurogenesis and skin development. Accordingly, we could speculate that MEDNIK, a novel neurocutaneous syndrome, is caused by impaired development of various neural networks in the spinal cord and in the brain, explaining the multifunctional human deficits such as ataxia, peripheral neuropathy, and mental retardation, which are concomitant with a perturbation in epithelial cell development in the skin and in the digestive system. We speculate that these generalized effects of AP1S1 Fig. 2. Evaluating a mutation found in human AP1S1 gene (MEDNIK syndrome) using zebrafish. (A) An illustration of the human AP1S1 gene to show the location of the mutation found in MEDNIK patients. An A to G mutation in the acceptor splice site of exon 3 resulted in skipping of this exon and led to a premature stop codon. The use of an alternative cryptic splice site located 9 bp downstream of the start of the 3rd exon resulted in mRNA lacking 9 bp. The location of the two different morpholinos targeting either ATG or exon 3 acceptor splice site of the zebrafish ap1s1 gene is indicated in green. The inset on the right depicts the clathrin adaptor protein 1 (AP-1) and its 4 subunits. It is the small subunit σ1 (red), which is mutated in MEDNIK patients. (B) Ap1s1 knockdown (KD) in zebrafish using morpholino oligo nucleotides targeting the N-terminal or the splice site (splice site, intron 2) resulted in a similar characteristic morphological phenotype (48 h post-fertilization), which is rescued by co-injection of the wild-type human AP1S1 (rescue + WT HmRNA) or the mRNA lacking 9 bp (rescue + −9 bp HmRNA), but not by the mutated human mRNA missing exon 3 (rescue + −exon3HmRNA). (C) Illustration of the predicated human AP1S1 proteins: normal protein containing 55AA (WT, black); truncated protein containing 19AA (red), the result of skipping exon 3; and a protein missing 3AA, a result of the alternative splicing (blue). Western blot analysis of Ap1s1 proteins from human (left) and zebrafish (right) to show that Ap1s1 protein is hardly expressed in MEDNIK patients and in ap1s1 knockdown (KD) zebrafish larva. To normalize the Western blot analysis, proteins extracted from WT, KD, and CTRL larvae were incubated with anti-actin. CTRL = control, WT = wild type. (D) Localization of Ap1s1 in skin cells of zebrafish wholemounts is illustrated using anti-Ap1s1 antibody immunofluorescence. Cell membrane (polygonal) and a well-defined perinuclear ring can be nicely observed in both normal (WT) and rescued larvae (rescue + WTHmRNA), whereas only a residual and diffuse staining could be observed in knockdown (KD) larvae (modified from Ref. [47]). E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 mutation are due to a widespread deficit in vesicular transport and protein sorting. 3. Degenerative spinal cord disorders Zebrafish have also been proven to be particularly effective as a model for motor dysfunction [52]. Our group has worked extensively on motor neuron diseases, including the most common of these disorders, amyotrophic lateral sclerosis (ALS). The advantage of studying motor neuron diseases in zebrafish consists in the rapid development of the spinal cord and allowing analysis of motor neuron branching patterns as early as 24 hpf. In addition, responses to touch and swimming can be monitored following hatching around 48 hpf [53]. SOD1 (ALS1) is known to cause ALS in 10–20% of familial ALS cases mainly through an autosomal dominant toxic gain of function and has 70% amino acid identity to the zebrafish homologue. Overexpression of mutant SOD1 in zebrafish leads to short motor axons with premature branching [54]. Interestingly, several other genes implicated in motoneuron degeneration show a similar phenotype when tested in zebrafish. Alsin (ALS2) is the gene mutated in juvenile ALS through autosomal recessive mode of action, which leads in most cases to protein truncation. Corresponding to this, in zebrafish embryos, knockdown of Alsin (61% homology) by AMO causes shortening of motor axons and also loss of neurons in the spinal cord [55]. ELP3 was identified through association studies performed in ALS patients and in a Drosophila screen for genes important for neuronal survival [56]. Knockdown of this gene (ELP3) in zebrafish caused increased branching and shortened motor axons [56]. Recently, TDP-43 was found enriched in inclusion bodies from spinal cord autopsy tissue obtained from ALS patients [57]. Moreover, about 30 mutations have been identified in a considerable number of ALS patients, suggesting that this protein plays an important role in disease pathogenesis [58,59]. The zebrafish ztardbp gene product (TDP-43) has 73% homology to the human protein. Knockdown of TDP-43 (tardbp) in zebrafish embryos led to motor neuron branching and motility deficits (Fig. 3). This AMO phenotype was rescued by coexpressing human wild-type (WT) TDP-43 (which is not targeted by the AMO) (Fig. 3). However, three ALS-related mutations of human TDP-43 identified in both in SALS and FALS cases (A315T, G348C, A382T) failed to rescue these phenotypes, indicating that these mutants are pathogenic (Fig. 3). A similar phenotype was observed when mutant (but not wt) TDP-43 was overexpressed, with the G348C mutation being the most penetrant (Fig. 3). These results indicate that TDP-43 mutations cause motor defects, suggesting that both a loss as well as a gain of function may be involved in the molecular mechanism of pathogenesis [125]. Finally, it is interesting to note that a number of groups generated ALS2 knockout mice yet failed to observe motor deficits associated with motor neuron degeneration and concluded that ALS2 was not an important gene for ALS [55,60]. However, the first three exons of ALS2 were left intact in these mice. We found that the motor phenotype in zebrafish upon ALS2 KD could be partially rescued upon overexpression of the alternative transcript of the first three exons of ALS2, indicating that complete disruption could in fact be pathogenic in ALS [55]. This nicely illustrates the usefulness of zebrafish knockdown models for more complete disruption of gene function in some circumstances. Zebrafish have been also widely used to study the functional role of gene mutations in another motor neuron disease, hereditary spastic paraplegia (HSP). Over 30 loci are known for this disorder, and at least 20 genes have been identified to cause spastic paraplegia [61]. Our group identified missense variants in the KIAA0196 gene at the SPG8 locus and validated these mutations using the zebrafish model (87% homology to the human protein) [62]. Knockdown of the zebrafish homolog of KIAA0196 gene caused a curly-tail phenotype coupled with shorter motor axons. This phenotype could be partially rescued by the human WT KIAA0196 but not the two mutants identified in 339 HSP patients [62]. Recently, SLC33A1 was identified as the gene responsible for SPG42 [63]. Knockdown of this gene in zebrafish embryos using AMO caused a curly-tail phenotype and defective axon outgrowth from the spinal cord [63]. Finally, knockdown of spastin (SPG4) caused widespread defects in neuronal connectivity and extensive CNS-specific apoptosis [64]. Spinal muscular atrophy (SMA) is an autosomally recessive motor neuron disorder caused by mutations in the survival motor neuron gene (SMN1). In a series of publications, Beattie's laboratory has demonstrated that knockdown of the SMN1 in zebrafish embryos (49% homology to human protein) causes motor axon outgrowth and pathfinding defects in early development [52,65]. Further, in a series of elegant rescue experiments, this group was able to show that a conserved region in exon 7 of the SMN1 gene (QNQKE) is critical for axonal outgrowth and that the plastin 3 (PLS3) may be important for axonal outgrowth since overexpression of plastin 3 rescued the phenotype caused by SMN1 knockdown [66,67]. Finally, transgenic expression of homozygous deletion found in SMA patients that leads to truncated SMN1 protein as well as knocking out the Smn1 gene in zebrafish caused deficits in the neuromuscular junction formation and death at the larval stage [68]. 4. Developmental disorders Although this developmental model is proving useful in the study of late-onset degenerative diseases, zebrafish are just starting to be used in the study of developmental brain diseases. Interesting speculations have been made on the potential of zebrafish for modeling developmental psychiatric disorders such as autism [69], but the lag in advancing these models is perhaps because the clinical phenotypes are so subtle and specific to humans suffering from disorders such as autism, schizophrenia, non-syndromic mental retardation, and others that zebrafish may appear as a doubtful model. So far (and very recently), the only genes linked to schizophrenia to be studied in zebrafish using AMO methods are DISC1 and NRG1 [70]. The receptor tyrosine kinase Met has been implicated in cerebellar development and autism, and knockdown of either Met alone or both of its Hgf ligands (two genes in zebrafish) together using ATG or splice junction AMOs results in abnormal cerebellar and facial motoneuron development [71]. However, neither the human homologs of these genes nor their disease-related mutations have been tested in zebrafish. Our new ongoing genomics approach, based on large-scale re-sequencing of human synaptic genes in patients suffering from some of these disorders, is identifying mutations that can be validated, if not exactly modeled, in zebrafish. As a part of our “Synapse to Disease” project, we are using a combination of AMO knockdown and gene overexpression to functionally validate novel synaptic gene mutations identified in large cohorts of patients with autism, schizophrenia, or nonsyndromic mental retardation. Hundreds of synaptic genes were selected based on published studies and databases. Many genes were chosen because they are X-linked [72] as there are many more affected males than females in autism [73]. Other genes that were selected are those of the glutamate receptor complex because of their association to neurodevelopmental diseases [74]. So far, the exonic regions of over 400 genes have been sequenced, and 15 de novo mutations (present in the patients and not their parents) were genetically validated (manuscript submitted). We have identified deleterious rare de novo mutations in Shank3, IL1RAPL1, NRXN1, and KIF17 genes, and all of these genes have orthologs in zebrafish. We have identified a de novo mutation in the Shank3 gene in a patient with autism [75] as Shank3 deletions or duplications were previously identified in patients with autism [76]. Shank3 encodes a scaffolding protein found in excitatory synapses directly opposite to the presynaptic active zone. The Shank proteins link ionotropic and 340 E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 Fig. 3. Motor phenotype of gain and loss of function of TDP-43 in zebrafish. (A) A schematic representation of TDP-43 protein showing the functional domains, localization of FLAG (3 kDa) and myc tags (12 kDa), and sites where the G348C mutation found in ALS patients was introduced by site-directed mutagenesis. (B) Locomotor phenotype. Zebrafish embryos develop a touch-evoked escape response as seen when embryos are injected with WT human TDP-43 RNA. Expression of TDP-43 G348C RNA causes a deficiency in the touch-evoked response. A similar phenotype is observed when the zf tdp-43 expression was knocked down using a specific AMO. This phenotype was rescued with expression of WT but not G348C RNA. (C) Motor axonal deficits. The behavioral (motor) phenotype was selectively associated with a shortening of the axon and premature branching in motor neurons (see G348C TDP-43 and tdp-43AMO) (modified from Kabashi et al., accepted at Hum. Mol. Genet.). metabotropic glutamate receptor complexes together and to the cytoskeleton to regulate the structural organization of dentritic spines. The splice mutation identified in the Shank3 gene would result in a truncated protein lacking the Homer, Cortactin, and SAM domains important for spine induction and dentritic targeting [77]. Zebrafish have two Shank3 homologous genes (shank3a and shank3b). The identity to human protein is overall 65% and 62% for shank3a and shank3b, respectively, but when we compare every individual functional domain to the human one, the amino acids identity reach 80% and is over 90% for the PDZ domain. The phenotype of Shank3 knockdown zebrafish was not reported in the literature yet, and we observed major motility deficits that could be rescued by the rat mRNA but not by some of the mutations [126]. We also found de novo mutations in IL1RAPL1 in autistic patients [78]. IL1RAPL1 and 2 are plasma membrane proteins that belong to a class of the interleukin-1 receptor family characterized by a 150-aa Cterminal domain that interacts with Neuronal Calcium Sensor-1 (NCS1) [79,80]. A recent study from [81] used IL1RAPL1-deficient mouse to show that IL1RAPL1 controls inhibitory network during cerebellar development. In one patient, a de novo frameshift mutation in IL1RAPL1 causes a premature stop of the translation (I367fs) resulting in a truncated protein, lacking the intracellular TIR domain and the Cterminal domain interacting with NCS-1 [78]. Moreover, a large deletion of exons 3 to 7 of IL1RAPL1 in three brothers with autism and/or non-syndromic mental retardation was also identified in our study. Mutations resulting in deletion of the TIR and the C-terminal domains were previously identified in patients with non-syndromic mental retardation [82,83], suggesting that those domains are important to the protein function. Two genes encode IL1RAPL1 in zebrafish (il1rapl1a and il1rapl1b) and share 69% and 75% identity, respectively, with human protein. The divergence is mainly in the signal peptide (il1rapl1a: 42%, b: 47%). Recently, a study using zebrafish il1rapl1a showed that the mRNA is constantly expressed during early embryonic development, indicating that it is maternally provided [84]. In this study, they also showed that knockdown of the other gene, il1rapl1b, using an ATG AMO caused a selective loss of synaptic endings in olfactory neurons and that il1rapl1b C terminus and TIR domains regulate the synaptic vesicles accumulation and morphological remodeling of axon terminals during synapse formation [84]. These observations suggest that il1rapl1b may be essential for synapse formation and indicate how a loss-of-function IL1RAPL1 mutant could affect brain neurodevelopment and its possible association to autism and non-syndromic mental retardation. To determine more specifically the function of IL1RAPL1 in the context of schizophrenia, we knocked down the expression of zebrafish il1rapl1a with an ATG AMO and observed a severe phenotype with incomplete development of the embryo, which could not be rescued by human mRNA. These preliminary (unpublished) results indicate that IL1RAPL1 plays an important function during early embryonic development. Further work, such as with splice junction AMOs that E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 spare maternal transcripts that are essential for early development, is required in order to develop a pathogenic validation. Neurexins are predominantly pre-synaptic cell-adhesion molecules. They can induce pre-synaptic differentiation by interacting with neuroligins. There are three neurexin genes (NRXN1, 2, and 3), each of which encodes two major variants (alpha and beta). Only NRXN1 gene was shown to be disrupted using CNV analysis in autism [85,86]. More recently, NRXN1 gene has been involved in CNV found in SCZ patients [87,88]. We have identified an insertion of 4 bp predicted to cause a frameshift with a premature stop affecting the two major variants alpha and beta of NRXN1 in SCZ patient. The protein is predicted to miss the transmembrane and intracellular domains. Two orthologs of NRXN1 have been identified in zebrafish (Nrxn1a and 1b) with an identity to the human protein over 70%. The greatest variability is found in the signal peptide (Nrxn1a, 27%; 1b 32%). An expression analysis showed that all three Nrxn genes are expressed during zebrafish embryonic development and some specific isoforms of Nrxn1a expressed at different stage of the development [89]. These results indicate the potential for developing zebrafish models of neurexin mutations. However, the fact that there exist thousands of isoforms of Nrxn1 and the amino acids of the signal peptide are divergent could make the functional validation in zebrafish more challenging. Kinesin 17 (KIF17) is a member of the kinesin superfamily containing a motor domain for ATP hydrolysis. KIF17 binds to the scafolding Mint1 to transport the NMDA receptor subunit 2B (NR2B) to the dendrites along microtubules [90,91]. In addition, the K+ channel Kv4.2, a major regulator of dendritic excitability, is also transported to the dentrites by KIF17 [92]. In mice, overexpression of Kif17 enhances spatial and working memory [93]. The expression of KIF17 was shown to be altered in a mouse models for Down syndrome (trisomy 21) leading to learning deficit [94]. We have identified a nonsense mutation resulting in a protein that lacks the tail domain in a SCZ patient. This domain was shown to interact with Mint1 following its phosphorylation by CaMKII [95]. One orthologous sequence was identified in zebrafish with an identity of 61% compare to the human amino acids. The divergence was mainly observed in the motifs important for cargo binding. The zebrafish KIF17 is widely expressed in the nervous system and retina [96]. By using low doses splicing AMO, this group demonstrates that KIF17 is essential for vertebrate photoreceptor development and seems to have little effect on the global zebrafish development. At higher doses of either an ATG or a splice junction AMO that micked the effect of the de novo nonsense truncation mutation in schizophrenia, we observed a characteristic dose-dependent morphological trait with stunted, curly embryos [127]. This validates that the nonsense mutation is a loss-of-function pathogenic effect. Together, these results with several zebrafish developmental genes indicate the potential usefulness of this model for validating brain disease mutations, but clearly each gene requires its own set of carefully designed experiments. Whereas some human genes can be more difficult to study in zebrafish, with others, one can observe faithful disease phenotypes that can be characterized in detail such as for AP1S1 in MEDNIK and ALS2 in ALS and thus provide novel insights to these disorders. We have observed some knockdown phenotypes in zebrafish that are not as obvious in knockout mice, which may reflect a lower degree of compensation in zebrafish, perhaps because their reproductive biology is based more on the quantity of offspring rather than necessarily on the quality of developmental compensation. 5. Future directions Several approaches have the potential of further advancing zebrafish models of human CNS disorders and making them uniquely useful, in particular targeted expression, multigenic analysis and chemical genetics. 341 5.1. Targeted expression Most of the approaches taken to date to express foreign genes in zebrafish are based on injection of mRNA for transient expression throughout the embryo. This simple approach is valid for ubiquitously and early expressed genes such as AP1S1, SOD1, TARDBP and members of the ubiquination and beta-secretase complexes discussed above. However, mRNA injection risks expressing genes ectopically, such as in the case of synaptic genes, and is usually only efficient at embryonic stages, prior to mRNA degradation that occurs within a few days. The latter thus limits the time course for phenotypic analysis, which is particularly problematic for studies of neurogeneration in (often) adult-onset human diseases. More specific expression patterns can be achieved by stable transgenesis using the efficient Tol2 transposon system [97], but this is best suited for generating long-term models (rather than for preliminary or large-scale screens) as it is a slower process that requires the raising of founder fish and their out-crossing to verify stable and specific transgenic lines. A number of promoters are available for neuralspecific transgenic expression. For example, the α-tubulin promoter [98] was the first used to drive pan-neuronal expression and, as described above, the HuC promoter expressed in post-mitotic neurons [19] and a neural-specific variant of the GATA2 promoter [18] are also effective at early stages. More selective subsets of neurons can be targeted using more restrictive promoters such as the dopaminergic-specific vesicular monoamine transporter 2 promoter [31]. In the spinal cord, which is an important region for studies of motor dysfunction and synaptic transmission, the developmentally conserved transcriptional code [99] has permitted the generation of transgenic lines for selective types of neurons. These include motoneurons: HB9 [100], Isl-1 [101], commissural neurons: Evx1 [102], mostly glycinergic neurons: Pax2.1 [103]; vGlyT2 [104], mostly glutamatergic neurons: Alx/Vsx2 [105], as well as sensory neurons: Ngn1 [106] and oligodendrocytes: Nkx2.2a [107], olig2 [108] and sox10 [109]. However, these stable transgenic models also have their limitations, and here, we consider three major ones that can be at least partially avoided depending on necessity: ectopic expression, toxic expression, and genetic background. Although promoter-based stable transgenic lines confer more selective expression patterns, rarely do they perfectly recapitulate the natural expression patterns as often only subsets of enhancer elements are used in generating the transgenic constructs. Indeed, this is a limitation of the lack of a knock-in technology in zebrafish, requiring the integration of novel constructs in a wild-type background (considered below). However, perfectly faithful expression patterns are not always essential as expression in incomplete or expanded subsets of neurons can be useful to study. An approach that provides a more accurate expression pattern is the use of bacterial artificial chromosomes (BACs), such as with the amyloid line mentioned above [21]. In principle (although we are unaware of its practice with zebrafish), humanized models could be generated based on transgenesis with human BACs, as is commonly done with mice. However, vertebrate genes can be hundreds of kilobases in size and are not always completely contained within single BACs. As an alternative, pufferfish (Takifugu rubripes) BACs can be used as the intergenic regions are much smaller, an approach we have taken in generating an Evx1 transgenic line [102], although even in this case less than 80% of the neurons labeled by a selective antibody to evx1 also expressed GFP, indicating that not all evx1 cells transgenically express GFP. A second issue in considering stable transgenic lines is toxicity, as many of the transgenes bearing disease-related mutations can be embryonic-lethal. A simple approach is to drive expression from a heat shock promoter [110], although this may lose in generalized spatial expression what it gains in restricted temporal activation. A powerful alternative is the use of combinatorial expression systems, 342 E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 such as the cre-loxP system developed in mice and the yeast Gal4-UAS binary system popular for invertebrate genetics, both of which have found applications in zebrafish [111,112]. Examples are the Evx1-Gal4 [102] and HuC-Gal4 [19] lines described earlier. The latter drives transcription in both directions from the UAS in order to express two constructs simultaneoulsy, such as a gene of interest (in this case human P301L tau) and a reporter (dsRed) in the same cells. Another technique for expressing more than one gene from the same construct is to use the self-cleaving viral 2A peptide as a linker between the genes of interest [113]. A final concern is the creation of transgenic lines in a genetic background containing the endogenous gene as (in the absence of knock-in technology) this results in the addition of another gene. One possibility that is rarely used is to create a transgenic in a knockout background. In principle, this can be done if mutants have been isolated in screens for related phenotypes or by TILLING (Targeted Induced Local Lesions in Genomes) [114] available through a new consortium (see https://webapps.fhcrc.org/science/tilling/). A new technology is the targeted lesioning of genes by zinc finger nucleases (ZFN) [115], although this is a complex and expensive technology that at the moment is beyond the reach of small laboratories. Preliminary success in targeting genes in zebrafish is encouraging, and the possibility to use homologous recombination, rather than nonhomologous end joining, to repair ZFN-induced double-strand breaks may permit the development of a knock-in approach for zebrafish [116]. Much works remain to be done before this enticing possibility becomes practical, but nonetheless, a number of transgenic approaches are now available for refined studies of gene expression and function in zebrafish, including human genes bearing diseaserelated mutations. 5.2. Multigenic analysis Although highly penetrant mutations of single genes are often implicated in CNS disorders, such as the majority considered in this review, most diseases are multigenic in nature, requiring multiple weakly penetrating mutations in predisposing genetic backgrounds in order to produce the disease. Existing genetic models most easily address single gene mutations (or rather have focused on these simple cases), but clearly, the future of these models lies in their ability to provide insights to the nature of genetic interactions in disease. The zebrafish may prove to be particularly amenable to multigenic analysis by a combination of the approaches described above. The use of two or even three AMOs is a simple solution once each AMO is validated. For example, in the context of Parkinson's disease, simultaneous knockdown of DJ-1 and p53 was found to cause dopaminergic neuronal loss [28] and knockdown of both Hgf ligands was necessary to induce Met-related cerebellar defects [71]. Also, photoactivatable caged AMOs can be used to release these at specific time points [117]. This approach seems promising, but unfortunately, caged AMOs are not available commercially yet. Another major limitation of the AMO is the necessity to inject in embryo at 1 or 2 cells stage. An interesting alternative would be to use the vivo-AMOs that allow making gene knockdown and splice modification in adult animals. The vivo-AMOs are modified AMO able to enter cells by endocytosis. They could be conjugated with a dendrimeric octaguanidine (Gene Tools) or a cell-penetrating peptide linked to phosphorodiamidate (PPMO) (AVI BioPharma Inc.). Vivo-AMOs were shown to be effective in mice [118,119], and they are currently tested in adult zebrafish [3]. By performing AMO injections in mutants or transgenics, either stable or inducible, it should in principle be possible to readily test several genes simultaneously or in different combinations, bestowing zebrafish with a polyvalence that is difficult to envisage for other models such as mice. In the case of mutant and transgenic backgrounds, comparative microarray screening [120] becomes feasible, particularly as multiple conditions (e.g., different mutant alleles) permit internal verification and comparison of the array results. A pertinent example is the recent microarray analysis [121] of three alleles of the neurogenic mindbomb mutant affecting Notch signaling for which each allele affected the expression of hundreds of genes but less than 100 were commonly affected. Thus, zebrafish have a unique potential for rapidly analyzing multiple genes and dissecting complex genetic pathways. 5.3. Chemical genetics While zebrafish retain advantages for studying the genetics, biology, and pathobiology of disease genes of interest, they also have the major advantage of being the only vertebrate model amenable to large chemical genetic screens [122]. Small chemical libraries of over 10,000 molecules, including approved drugs and randomly synthesized organic molecules of unknown function, have been screened in zebrafish models of cancer [123] and cardiovascular disorders [124], and some of these are in clinical testing. With the numerous models of brain diseases being developed in zebrafish, this approach holds much potential for drug discovery. This is particularly attractive in zebrafish as functional chemical genetic screens in vivo with living embryos offer the potential of discovering therapeutics without prior knowledge or need for rational screening designs. Thus, zebrafish are a unique model ranging from molecular genetic validation to drug discovery with important applications on the horizon for brain diseases, including many complex and untreatable disorders. References [1] J.N. Volff, Genome evolution and biodiversity in teleost fish, Heredity 94 (2005) 280–294. [2] A. Nasevicius, S.C. Ekker, Effective targeted gene ‘knockdown’ in zebrafish, Nat. Genet. 26 (2000) 216–220. [3] J.D. Moulton, Y.L. Yan, Using morpholinos to control gene expression, Curr. Protoc. Mol. Biol. (2008) Chapter 26, Unit 26 28. [4] D.L. Turner, H. Weintraub, Expression of achaete-scute homolog 3 in Xenopus embryos converts ectodermal cells to a neural fate, Genes Dev. 8 (1994) 1434–1447. [5] H. Ro, K. Soun, E.J. Kim, M. Rhee, Novel vector systems optimized for injecting in vitro-synthesized mRNA into zebrafish embryos, Mol. Cells 17 (2004) 373–376. [6] J.D. Best, W.K. Alderton, Zebrafish: an in vivo model for the study of neurological diseases, Neuropsychiatr. Dis. Treat. 4 (2008) 567–576. [7] P.W. Ingham, The power of the zebrafish for disease analysis, Hum. Mol. Genet. 18 (2009) R107–R112. [8] U. Leimer, K. Lun, H. Romig, J. Walter, J. Grunberg, M. Brand, C. Haass, Zebrafish (Danio rerio) presenilin promotes aberrant amyloid beta-peptide production and requires a critical aspartate residue for its function in amyloidogenesis, Biochemistry 38 (1999) 13602–13609. [9] S. Nornes, M. Newman, G. Verdile, S. Wells, C.L. Stoick-Cooper, B. Tucker, I. Frederich-Sleptsova, R. Martins, M. Lardelli, Interference with splicing of presenilin transcripts has potent dominant negative effects on presenilin activity, Hum. Mol. Genet. 17 (2008) 402–412. [10] M. Newman, B. Tucker, S. Nornes, A. Ward, M. Lardelli, Altering presenilin gene activity in zebrafish embryos causes changes in expression of genes with potential involvement in Alzheimer's disease pathogenesis, J. Alzheimers Dis. 16 (2009) 133–147. [11] C. Groth, S. Nornes, R. McCarty, R. Tamme, M. Lardelli, Identification of a second presenilin gene in zebrafish with similarity to the human Alzheimer's disease gene presenilin2, Dev. Genes Evol. 212 (2002) 486–490. [12] S. Nornes, C. Groth, E. Camp, P. Ey, M. Lardelli, Developmental control of presenilin1 expression, endoproteolysis, and interaction in zebrafish embryos, Exp. Cell Res. 289 (2003) 124–132. [13] A. Musa, H. Lehrach, V.A. Russo, Distinct expression patterns of two zebrafish homologues of the human APP gene during embryonic development, Dev. Genes Evol. 211 (2001) 563–567. [14] P. Joshi, J.O. Liang, K. DiMonte, J. Sullivan, S.W. Pimplikar, Amyloid precursor protein is required for convergent-extension movements during zebrafish development, Dev. Biol. 335 (2009) 1–11. [15] M. Newman, I.F. Musgrave, M. Lardelli, Alzheimer disease: amyloidogenesis, the presenilins and animal models, Biochim. Biophys. Acta 1772 (2007) 285–297. [16] A. Geling, H. Steiner, M. Willem, L. Bally-Cuif, C. Haass, A gamma-secretase inhibitor blocks Notch signaling in vivo and causes a severe neurogenic phenotype in zebrafish, EMBO Rep. 3 (2002) 688–694. [17] W.A. Campbell, H. Yang, H. Zetterberg, S. Baulac, J.A. Sears, T. Liu, S.T. Wong, T.P. Zhong, W. Xia, Zebrafish lacking Alzheimer presenilin enhancer 2 (Pen-2) E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] demonstrate excessive p53-dependent apoptosis and neuronal loss, J. Neurochem. 96 (2006) 1423–1440. H.G. Tomasiewicz, D.B. Flaherty, J.P. Soria, J.G. Wood, Transgenic zebrafish model of neurodegeneration, J. Neurosci. Res. 70 (2002) 734–745. D. Paquet, R. Bhat, A. Sydow, E.M. Mandelkow, S. Berg, S. Hellberg, J. Falting, M. Distel, R.W. Koster, B. Schmid, C. Haass, A zebrafish model of tauopathy allows in vivo imaging of neuronal cell death and drug evaluation, J. Clin. Invest. 119 (2009) 1382–1395. J.A. Lee, G.J. Cole, Generation of transgenic zebrafish expressing green fluorescent protein under control of zebrafish amyloid precursor protein gene regulatory elements, Zebrafish 4 (2007) 277–286. L.A. Shakes, T.L. Malcolm, K.L. Allen, S. De, K.R. Harewood, P.K. Chatterjee, Context dependent function of APPb enhancer identified using enhancer trapcontaining BACs as transgenes in zebrafish, Nucleic Acids Res. 36 (2008) 6237–6248. L. Flinn, S. Bretaud, C. Lo, P.W. Ingham, O. Bandmann, Zebrafish as a new animal model for movement disorders, J. Neurochem. 106 (2008) 1991–1997. S. Bretaud, S. Lee, S. Guo, Sensitivity of zebrafish to environmental toxins implicated in Parkinson's disease, Neurotoxicol. Teratol. 26 (2004) 857–864. C.S. Lam, V. Korzh, U. Strahle, Zebrafish embryos are susceptible to the dopaminergic neurotoxin MPTP, Eur. J. Neurosci. 21 (2005) 1758–1762. E.T. McKinley, T.C. Baranowski, D.O. Blavo, C. Cato, T.N. Doan, A.L. Rubinstein, Neuroprotection of MPTP-induced toxicity in zebrafish dopaminergic neurons, Brain Res. Mol. Brain Res. 141 (2005) 128–137. A. Abeliovich, M. Flint Beal, Parkinsonism genes: culprits and clues, J. Neurochem. 99 (2006) 1062–1072. O.L. Son, H.T. Kim, M.H. Ji, K.W. Yoo, M. Rhee, C.H. Kim, Cloning and expression analysis of a Parkinson's disease gene, uch-L1, and its promoter in zebrafish, Biochem. Biophys. Res. Commun. 312 (2003) 601–607. S. Bretaud, C. Allen, P.W. Ingham, O. Bandmann, p53-dependent neuronal cell death in a DJ-1-deficient zebrafish model of Parkinson's disease, J. Neurochem. 100 (2007) 1626–1635. L. Flinn, H. Mortiboys, K. Volkmann, R.W. Koster, P.W. Ingham, O. Bandmann, Complex I deficiency and dopaminergic neuronal cell loss in Parkin-deficient zebrafish (Danio rerio), Brain 132 (2009) 1613–1623. W. Boehmler, J. Petko, M. Woll, C. Frey, B. Thisse, C. Thisse, V.A. Canfield, R. Levenson, Identification of zebrafish A2 adenosine receptors and expression in developing embryos, Gene Expr. Patterns 9 (2009) 144–151. L. Wen, W. Wei, W. Gu, P. Huang, X. Ren, Z. Zhang, Z. Zhu, S. Lin, B. Zhang, Visualization of monoaminergic neurons and neurotoxicity of MPTP in live transgenic zebrafish, Dev. Biol. 314 (2008) 84–92. C.A. Karlovich, R.M. John, L. Ramirez, D.Y. Stainier, R.M. Myers, Characterization of the Huntington's disease (HD) gene homologue in the zebrafish Danio rerio, Gene 217 (1998) 117–125. A.L. Lumsden, T.L. Henshall, S. Dayan, M.T. Lardelli, R.I. Richards, Huntingtindeficient zebrafish exhibit defects in iron utilization and development, Hum. Mol. Genet. 16 (2007) 1905–1920. H. Diekmann, O. Anichtchik, A. Fleming, M. Futter, P. Goldsmith, A. Roach, D.C. Rubinsztein, Decreased BDNF levels are a major contributor to the embryonic phenotype of huntingtin knockdown zebrafish, J. Neurosci. 29 (2009) 1343–1349. N.W. Schiffer, S.A. Broadley, T. Hirschberger, P. Tavan, H.A. Kretzschmar, A. Giese, C. Haass, F.U. Hartl, B. Schmid, Identification of anti-prion compounds as efficient inhibitors of polyglutamine protein aggregation in a zebrafish model, J. Biol. Chem. 282 (2007) 9195–9203. V.M. Miller, R.F. Nelson, C.M. Gouvion, A. Williams, E. Rodriguez-Lebron, S.Q. Harper, B.L. Davidson, M.R. Rebagliati, H.L. Paulson, CHIP suppresses polyglutamine aggregation and toxicity in vitro and in vivo, J. Neurosci. 25 (2005) 9152–9161. A. Williams, S. Sarkar, P. Cuddon, E.K. Ttofi, S. Saiki, F.H. Siddiqi, L. Jahreiss, A. Fleming, D. Pask, P. Goldsmith, C.J. O'Kane, R.A. Floto, D.C. Rubinsztein, Novel targets for Huntington's disease in an mTOR-independent autophagy pathway, Nat. Chem. Biol. 4 (2008) 295–305. S.C. Baraban, Emerging epilepsy models: insights from mice, flies, worms and fish, Curr. Opin. Neurol. 20 (2007) 164–168. J.A. Tiedeken, J.S. Ramsdell, DDT exposure of zebrafish embryos enhances seizure susceptibility: relationship to fetal p,p′-DDE burden and domoic acid exposure of California sea lions, Environ. Health Perspect. 117 (2009) 68–73. S.C. Baraban, M.R. Taylor, P.A. Castro, H. Baier, Pentylenetetrazole induced changes in zebrafish behavior, neural activity and c-fos expression, Neuroscience 131 (2005) 759–768. J.A. Tiedeken, J.S. Ramsdell, Embryonic exposure to domoic acid increases the susceptibility of zebrafish larvae to the chemical convulsant pentylenetetrazole, Environ. Health Perspect. 115 (2007) 1547–1552. J.A. Tiedeken, J.S. Ramsdell, A.F. Ramsdell, Developmental toxicity of domoic acid in zebrafish (Danio rerio), Neurotoxicol. Teratol. 27 (2005) 711–717. S.C. Baraban, M.T. Dinday, P.A. Castro, S. Chege, S. Guyenet, M.R. Taylor, A largescale mutagenesis screen to identify seizure-resistant zebrafish, Epilepsia 48 (2007) 1151–1157. M.J. Winter, W.S. Redfern, A.J. Hayfield, S.F. Owen, J.P. Valentin, T.H. Hutchinson, Validation of a larval zebrafish locomotor assay for assessing the seizure liability of early-stage development drugs, J. Pharmacol. Toxicol. Methods 57 (2008) 176–187. A.G. Bassuk, R.H. Wallace, A. Buhr, A.R. Buller, Z. Afawi, M. Shimojo, S. Miyata, S. Chen, P. Gonzalez-Alegre, H.L. Griesbach, S. Wu, M. Nashelsky, E.K. Vladar, D. Antic, P.J. Ferguson, S. Cirak, T. Voit, M.P. Scott, J.D. Axelrod, C. Gurnett, A.S. Daoud, S. Kivity, M.Y. Neufeld, A. Mazarib, R. Straussberg, S. Walid, A.D. Korczyn, [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] 343 D.C. Slusarski, S.F. Berkovic, H.I. El-Shanti, A homozygous mutation in human PRICKLE1 causes an autosomal-recessive progressive myoclonus epilepsy–ataxia syndrome, Am. J. Hum. Genet. 83 (2008) 572–581. T.G. Saba, A. Montpetit, A. Verner, P. Rioux, T.J. Hudson, R. Drouin, C.A. Drouin, An atypical form of erythrokeratodermia variabilis maps to chromosome 7q22, Hum. Genet. 116 (2005) 167–171. A. Montpetit, S. Cote, E. Brustein, C.A. Drouin, L. Lapointe, M. Boudreau, C. Meloche, R. Drouin, T.J. Hudson, P. Drapeau, P. Cossette, Disruption of AP1S1, causing a novel neurocutaneous syndrome, perturbs development of the skin and spinal cord, PLoS Genet. 4 (2008) e1000296. J.S. Bonifacino, B.S. Glick, The mechanisms of vesicle budding and fusion, Cell 116 (2004) 153–166. M.S. Robinson, Adaptable adaptors for coated vesicles, Trends Cell Biol. 14 (2004) 167–174. M. Boehm, J.S. Bonifacino, Adaptins: the final recount, Mol. Biol. Cell 12 (2001) 2907–2920. D.J. Owen, B.M. Collins, P.R. Evans, Adaptors for clathrin coats: structure and function, Annu. Rev. Cell Dev. Biol. 20 (2004) 153–191. C.E. Beattie, T.L. Carrel, M.L. McWhorter, Fishing for a mechanism: using zebrafish to understand spinal muscular atrophy, J. Child Neurol. 22 (2007) 995–1003. P. Drapeau, L. Saint-Amant, R.R. Buss, M. Chong, J.R. McDearmid, E. Brustein, Development of the locomotor network in zebrafish, Prog. Neurobiol. 68 (2002) 85–111. R. Lemmens, A. Van Hoecke, N. Hersmus, V. Geelen, I. D'Hollander, V. Thijs, L. Van Den Bosch, P. Carmeliet, W. Robberecht, Overexpression of mutant superoxide dismutase 1 causes a motor axonopathy in the zebrafish, Hum. Mol. Genet. 16 (2007) 2359–2365. F. Gros-Louis, J. Kriz, E. Kabashi, J. McDearmid, S. Millecamps, M. Urushitani, L. Lin, P. Dion, Q. Zhu, P. Drapeau, J.P. Julien, G.A. Rouleau, Als2 mRNA splicing variants detected in KO mice rescue severe motor dysfunction phenotype in Als2 knock-down zebrafish, Hum. Mol. Genet. 17 (2008) 2691–2702. C.L. Simpson, R. Lemmens, K. Miskiewicz, W.J. Broom, V.K. Hansen, P.W. van Vught, J.E. Landers, P. Sapp, L. Van Den Bosch, J. Knight, B.M. Neale, M.R. Turner, J.H. Veldink, R.A. Ophoff, V.B. Tripathi, A. Beleza, M.N. Shah, P. Proitsi, A. Van Hoecke, P. Carmeliet, H.R. Horvitz, P.N. Leigh, C.E. Shaw, L.H. van den Berg, P.C. Sham, J.F. Powell, P. Verstreken, R.H. Brown Jr., W. Robberecht, A. Al-Chalabi, Variants of the elongator protein 3 (ELP3) gene are associated with motor neuron degeneration, Hum. Mol. Genet. 18 (2009) 472–481. M. Neumann, D.M. Sampathu, L.K. Kwong, A.C. Truax, M.C. Micsenyi, T.T. Chou, J. Bruce, T. Schuck, M. Grossman, C.M. Clark, L.F. McCluskey, B.L. Miller, E. Masliah, I.R. Mackenzie, H. Feldman, W. Feiden, H.A. Kretzschmar, J.Q. Trojanowski, V.M. Lee, Ubiquitinated TDP-43 in frontotemporal lobar degeneration and amyotrophic lateral sclerosis, Science 314 (2006) 130–133. E. Kabashi, P.N. Valdmanis, P. Dion, D. Spiegelman, B.J. McConkey, C. Vande Velde, J.P. Bouchard, L. Lacomblez, K. Pochigaeva, F. Salachas, P.F. Pradat, W. Camu, V. Meininger, N. Dupre, G.A. Rouleau, TARDBP mutations in individuals with sporadic and familial amyotrophic lateral sclerosis, Nat. Genet. 40 (2008) 572–574. J. Sreedharan, I.P. Blair, V.B. Tripathi, X. Hu, C. Vance, B. Rogelj, S. Ackerley, J.C. Durnall, K.L. Williams, E. Buratti, F. Baralle, J. de Belleroche, J.D. Mitchell, P.N. Leigh, A. Al-Chalabi, C.C. Miller, G. Nicholson, C.E. Shaw, TDP-43 mutations in familial and sporadic amyotrophic lateral sclerosis, Science 319 (2008) 1668–1672. S. Hadano, S.C. Benn, S. Kakuta, A. Otomo, K. Sudo, R. Kunita, K. SuzukiUtsunomiya, H. Mizumura, J.M. Shefner, G.A. Cox, Y. Iwakura, R.H. Brown Jr., J.E. Ikeda, Mice deficient in the Rab5 guanine nucleotide exchange factor ALS2/alsin exhibit age-dependent neurological deficits and altered endosome trafficking, Hum. Mol. Genet. 15 (2006) 233–250. S. Salinas, C. Proukakis, A. Crosby, T.T. Warner, Hereditary spastic paraplegia: clinical features and pathogenetic mechanisms, Lancet Neurol. 7 (2008) 1127–1138. P.N. Valdmanis, I.A. Meijer, A. Reynolds, A. Lei, P. MacLeod, D. Schlesinger, M. Zatz, E. Reid, P.A. Dion, P. Drapeau, G.A. Rouleau, Mutations in the KIAA0196 gene at the SPG8 locus cause hereditary spastic paraplegia, Am. J. Hum. Genet. 80 (2007) 152–161. P. Lin, J. Li, Q. Liu, F. Mao, R. Qiu, H. Hu, Y. Song, Y. Yang, G. Gao, C. Yan, W. Yang, C. Shao, Y. Gong, A missense mutation in SLC33A1, which encodes the acetyl-CoA transporter, causes autosomal-dominant spastic paraplegia (SPG42), Am. J. Hum. Genet. 83 (2008) 752–759. J.D. Wood, J.A. Landers, M. Bingley, C.J. McDermott, V. Thomas-McArthur, L.J. Gleadall, P.J. Shaw, V.T. Cunliffe, The microtubule-severing protein Spastin is essential for axon outgrowth in the zebrafish embryo, Hum. Mol. Genet. 15 (2006) 2763–2771. M.L. McWhorter, U.R. Monani, A.H. Burghes, C.E. Beattie, Knockdown of the survival motor neuron (Smn) protein in zebrafish causes defects in motor axon outgrowth and pathfinding, J. Cell Biol. 162 (2003) 919–931. T.L. Carrel, M.L. McWhorter, E. Workman, H. Zhang, E.C. Wolstencroft, C. Lorson, G.J. Bassell, A.H. Burghes, C.E. Beattie, Survival motor neuron function in motor axons is independent of functions required for small nuclear ribonucleoprotein biogenesis, J. Neurosci. 26 (2006) 11014–11022. G.E. Oprea, S. Krober, M.L. McWhorter, W. Rossoll, S. Muller, M. Krawczak, G.J. Bassell, C.E. Beattie, B. Wirth, Plastin 3 is a protective modifier of autosomal recessive spinal muscular atrophy, Science 320 (2008) 524–527. K.L. Boon, S. Xiao, M.L. McWhorter, T. Donn, E. Wolf-Saxon, M.T. Bohnsack, C.B. Moens, C.E. Beattie, Zebrafish survival motor neuron mutants exhibit presynaptic neuromuscular junction defects, Hum. Mol. Genet. 18 (2009) 3615–3625. 344 E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 [69] V. Tropepe, H.L. Sive, Can zebrafish be used as a model to study the neurodevelopmental causes of autism? Genes Brain Behav. 2 (2003) 268–281. [70] J.D. Wood, F. Bonath, S. Kumar, C.A. Ross, V.T. Cunliffe, Disrupted-inschizophrenia 1 and neuregulin 1 are required for the specification of oligodendrocytes and neurones in the zebrafish brain, Hum. Mol. Genet. 18 (2009) 391–404. [71] G.E. Elsen, L.Y. Choi, V.E. Prince, R.K. Ho, The autism susceptibility gene met regulates zebrafish cerebellar development and facial motor neuron migration, Dev. Biol. 335 (2009) 78–92. [72] F. Laumonnier, P.C. Cuthbert, S.G. Grant, The role of neuronal complexes in human X-linked brain diseases, Am. J. Hum. Genet. 80 (2007) 205–220. [73] D.H. Skuse, Imprinting, the X-chromosome, and the male brain: explaining sex differences in the liability to autism, Pediatr. Res. 47 (2000) 9–16. [74] S.G. Grant, M.C. Marshall, K.L. Page, M.A. Cumiskey, J.D. Armstrong, Synapse proteomics of multiprotein complexes: en route from genes to nervous system diseases, Hum. Mol. Genet. 14 (Spec No. 2) (2005) R225–R234. [75] J. Gauthier, D. Spiegelman, A. Piton, R.G. Lafreniere, S. Laurent, J. St-Onge, L. Lapointe, F.F. Hamdan, P. Cossette, L. Mottron, E. Fombonne, R. Joober, C. Marineau, P. Drapeau, G.A. Rouleau, Novel de novo SHANK3 mutation in autistic patients, Am. J. Med. Genet. B Neuropsychiatr. Genet. 150B (2009) 421–424. [76] C.M. Durand, C. Betancur, T.M. Boeckers, J. Bockmann, P. Chaste, F. Fauchereau, G. Nygren, M. Rastam, I.C. Gillberg, H. Anckarsater, E. Sponheim, H. Goubran-Botros, R. Delorme, N. Chabane, M.C. Mouren-Simeoni, P. de Mas, E. Bieth, B. Roge, D. Heron, L. Burglen, C. Gillberg, M. Leboyer, T. Bourgeron, Mutations in the gene encoding the synaptic scaffolding protein SHANK3 are associated with autism spectrum disorders, Nat. Genet. 39 (2007) 25–27. [77] G. Roussignol, F. Ango, S. Romorini, J.C. Tu, C. Sala, P.F. Worley, J. Bockaert, L. Fagni, Shank expression is sufficient to induce functional dendritic spine synapses in aspiny neurons, J. Neurosci. 25 (2005) 3560–3570. [78] A. Piton, J.L. Michaud, H. Peng, S. Aradhya, J. Gauthier, L. Mottron, N. Champagne, R.G. Lafreniere, F.F. Hamdan, R. Joober, E. Fombonne, C. Marineau, P. Cossette, M.P. Dube, P. Haghighi, P. Drapeau, P.A. Barker, S. Carbonetto, G.A. Rouleau, Mutations in the calcium-related gene IL1RAPL1 are associated with autism, Hum. Mol. Genet. 17 (2008) 3965–3974. [79] N. Bahi, G. Friocourt, A. Carrie, M.E. Graham, J.L. Weiss, P. Chafey, F. Fauchereau, R.D. Burgoyne, J. Chelly, IL1 receptor accessory protein like, a protein involved in X-linked mental retardation, interacts with Neuronal Calcium Sensor-1 and regulates exocytosis, Hum. Mol. Genet. 12 (2003) 1415–1425. [80] F. Gambino, A. Pavlowsky, A. Begle, J.L. Dupont, N. Bahi, R. Courjaret, R. Gardette, H. Hadjkacem, H. Skala, B. Poulain, J. Chelly, N. Vitale, Y. Humeau, IL1-receptor accessory protein-like 1 (IL1RAPL1), a protein involved in cognitive functions, regulates N-type Ca2+-channel and neurite elongation, Proc. Natl. Acad. Sci. U. S. A. 104 (2007) 9063–9068. [81] F. Gambino, M. Kneib, A. Pavlowsky, H. Skala, S. Heitz, N. Vitale, B. Poulain, M. Khelfaoui, J. Chelly, P. Billuart, Y. Humeau, IL1RAPL1 controls inhibitory networks during cerebellar development in mice, Eur. J. Neurosci. 30 (2009) 1476–1486. [82] E. Tabolacci, M.G. Pomponi, R. Pietrobono, A. Terracciano, P. Chiurazzi, G. Neri, A truncating mutation in the IL1RAPL1 gene is responsible for X-linked mental retardation in the MRX21 family, Am. J. Med. Genet. 140 (2006) 482–487. [83] A. Carrie, L. Jun, T. Bienvenu, M.C. Vinet, N. McDonell, P. Couvert, R. Zemni, A. Cardona, G. Van Buggenhout, S. Frints, B. Hamel, C. Moraine, H.H. Ropers, T. Strom, G.R. Howell, A. Whittaker, M.T. Ross, A. Kahn, J.P. Fryns, C. Beldjord, P. Marynen, J. Chelly, A new member of the IL-1 receptor family highly expressed in hippocampus and involved in X-linked mental retardation, Nat. Genet. 23 (1999) 25–31. [84] T. Yoshida, M. Mishina, Zebrafish orthologue of mental retardation protein IL1RAPL1 regulates presynaptic differentiation, Mol. Cell. Neurosci. 39 (2008) 218–228. [85] P. Szatmari, A.D. Paterson, L. Zwaigenbaum, W. Roberts, J. Brian, X.Q. Liu, J.B. Vincent, J.L. Skaug, A.P. Thompson, L. Senman, L. Feuk, C. Qian, S.E. Bryson, M.B. Jones, C.R. Marshall, S.W. Scherer, V.J. Vieland, C. Bartlett, L.V. Mangin, R. Goedken, A. Segre, M.A. Pericak-Vance, M.L. Cuccaro, J.R. Gilbert, H.H. Wright, R.K. Abramson, C. Betancur, T. Bourgeron, C. Gillberg, M. Leboyer, J.D. Buxbaum, K.L. Davis, E. Hollander, J.M. Silverman, J. Hallmayer, L. Lotspeich, J.S. Sutcliffe, J.L. Haines, S.E. Folstein, J. Piven, T.H. Wassink, V. Sheffield, D.H. Geschwind, M. Bucan, W.T. Brown, R.M. Cantor, J.N. Constantino, T.C. Gilliam, M. Herbert, C. Lajonchere, D.H. Ledbetter, C. Lese-Martin, J. Miller, S. Nelson, C.A. Samango-Sprouse, S. Spence, M. State, R.E. Tanzi, H. Coon, G. Dawson, B. Devlin, A. Estes, P. Flodman, L. Klei, W.M. McMahon, N. Minshew, J. Munson, E. Korvatska, P.M. Rodier, G.D. Schellenberg, M. Smith, M.A. Spence, C. Stodgell, P.G. Tepper, E.M. Wijsman, C.E. Yu, B. Roge, C. Mantoulan, K. Wittemeyer, A. Poustka, B. Felder, S.M. Klauck, C. Schuster, F. Poustka, S. Bolte, S. Feineis-Matthews, E. Herbrecht, G. Schmotzer, J. Tsiantis, K. Papanikolaou, E. Maestrini, E. Bacchelli, F. Blasi, S. Carone, C. Toma, H. Van Engeland, M. de Jonge, C. Kemner, F. Koop, M. Langemeijer, C. Hijmans, W.G. Staal, G. Baird, P.F. Bolton, M.L. Rutter, E. Weisblatt, J. Green, C. Aldred, J.A. Wilkinson, A. Pickles, A. Le Couteur, T. Berney, H. McConachie, A.J. Bailey, K. Francis, G. Honeyman, A. Hutchinson, J.R. Parr, S. Wallace, A.P. Monaco, G. Barnby, K. Kobayashi, J.A. Lamb, I. Sousa, N. Sykes, E.H. Cook, S.J. Guter, B.L. Leventhal, J. Salt, C. Lord, C. Corsello, V. Hus, D.E. Weeks, F. Volkmar, M. Tauber, E. Fombonne, A. Shih, K.J. Meyer, Mapping autism risk loci using genetic linkage and chromosomal rearrangements, Nat. Genet. 39 (2007) 319–328. [86] H.G. Kim, S. Kishikawa, A.W. Higgins, I.S. Seong, D.J. Donovan, Y. Shen, E. Lally, L.A. Weiss, J. Najm, K. Kutsche, M. Descartes, L. Holt, S. Braddock, R. Troxell, L. Kaplan, F. Volkmar, A. Klin, K. Tsatsanis, D.J. Harris, I. Noens, D.L. Pauls, M.J. Daly, M.E. MacDonald, C.C. Morton, B.J. Quade, J.F. Gusella, Disruption of neurexin 1 associated with autism spectrum disorder, Am. J. Hum. Genet. 82 (2008) 199–207. [87] G. Kirov, D. Gumus, W. Chen, N. Norton, L. Georgieva, M. Sari, M.C. O'Donovan, F. Erdogan, M.J. Owen, H.H. Ropers, R. Ullmann, Comparative genome hybridization suggests a role for NRXN1 and APBA2 in schizophrenia, Hum. Mol. Genet. 17 (2008) 458–465. [88] D. Rujescu, A. Ingason, S. Cichon, O.P. Pietilainen, M.R. Barnes, T. Toulopoulou, M. Picchioni, E. Vassos, U. Ettinger, E. Bramon, R. Murray, M. Ruggeri, S. Tosato, C. Bonetto, S. Steinberg, E. Sigurdsson, T. Sigmundsson, H. Petursson, A. Gylfason, P.I. Olason, G. Hardarsson, G.A. Jonsdottir, O. Gustafsson, R. Fossdal, I. Giegling, H.J. Moller, A.M. Hartmann, P. Hoffmann, C. Crombie, G. Fraser, N. Walker, J. Lonnqvist, J. Suvisaari, A. Tuulio-Henriksson, S. Djurovic, I. Melle, O.A. Andreassen, T. Hansen, T. Werge, L.A. Kiemeney, B. Franke, J. Veltman, J.E. Buizer-Voskamp, C. Sabatti, R.A. Ophoff, M. Rietschel, M.M. Nothen, K. Stefansson, L. Peltonen, D. St Clair, H. Stefansson, D.A. Collier, Disruption of the neurexin 1 gene is associated with schizophrenia, Hum. Mol. Genet. 18 (2009) 988–996. [89] A. Rissone, M. Monopoli, M. Beltrame, F. Bussolino, F. Cotelli, M. Arese, Comparative genome analysis of the neurexin gene family in Danio rerio: insights into their functions and evolution, Mol. Biol. Evol. 24 (2007) 236–252. [90] M. Setou, T. Nakagawa, D.H. Seog, N. Hirokawa, Kinesin superfamily motor protein KIF17 and mLin-10 in NMDA receptor-containing vesicle transport, Science (New York, N.Y.) 288 (2000) 1796–1802. [91] L. Guillaud, M. Setou, N. Hirokawa, KIF17 dynamics and regulation of NR2B trafficking in hippocampal neurons, J. Neurosci. 23 (2003) 131–140. [92] P.J. Chu, J.F. Rivera, D.B. Arnold, A role for Kif17 in transport of Kv4.2, J. Biol. Chem. 281 (2006) 365–373. [93] R.W. Wong, M. Setou, J. Teng, Y. Takei, N. Hirokawa, Overexpression of motor protein KIF17 enhances spatial and working memory in transgenic mice, Proc. Natl. Acad. Sci. U. S. A. 99 (2002) 14500–14505. [94] R. Roberson, L. Toso, D. Abebe, C.Y. Spong, Altered expression of KIF17, a kinesin motor protein associated with NR2B trafficking, may mediate learning deficits in a Down syndrome mouse model, Am. J. Obstet. Gynecol. 198 (2008) 313 e311-314. [95] L. Guillaud, R. Wong, N. Hirokawa, Disruption of KIF17-Mint1 interaction by CaMKII-dependent phosphorylation: a molecular model of kinesin-cargo release, Nat. Cell Biol. 10 (2008) 19–29. [96] C. Insinna, N. Pathak, B. Perkins, I. Drummond, J.C. Besharse, The homodimeric kinesin, Kif17, is essential for vertebrate photoreceptor sensory outer segment development, Dev. Biol. 316 (2008) 160–170. [97] M.L. Suster, H. Kikuta, A. Urasaki, K. Asakawa, K. Kawakami, Transgenesis in zebrafish with the tol2 transposon system, Methods Mol. Biol. 561 (2009) 41–63. [98] D. Goldman, M. Hankin, Z. Li, X. Dai, J. Ding, Transgenic zebrafish for studying nervous system development and regeneration, Transgenic Res. 10 (2001) 21–33. [99] T.M. Jessell, Neuronal specification in the spinal cord: inductive signals and transcriptional codes, Nat. Rev. Genet. 1 (2000) 20–29. [100] T. Nakano, M. Windrem, V. Zappavigna, S.A. Goldman, Identification of a conserved 125 base-pair Hb9 enhancer that specifies gene expression to spinal motor neurons, Dev. Biol. 283 (2005) 474–485. [101] S. Higashijima, Y. Hotta, H. Okamoto, Visualization of cranial motor neurons in live transgenic zebrafish expressing green fluorescent protein under the control of the islet-1 promoter/enhancer, J. Neurosci. 20 (2000) 206–218. [102] M.L. Suster, A. Kania, M. Liao, K. Asakawa, F. Charron, K. Kawakami, P. Drapeau, A novel conserved evx1 enhancer links spinal interneuron morphology and cisregulation from fish to mammals, Dev. Biol. 325 (2009) 422–433. [103] A. Picker, S. Scholpp, H. Bohli, H. Takeda, M. Brand, A novel positive transcriptional feedback loop in midbrain–hindbrain boundary development is revealed through analysis of the zebrafish pax2.1 promoter in transgenic lines, Development 129 (2002) 3227–3239. [104] D.L. McLean, J. Fan, S. Higashijima, M.E. Hale, J.R. Fetcho, A topographic map of recruitment in spinal cord, Nature 446 (2007) 71–75. [105] Y. Kimura, Y. Okamura, S. Higashijima, alx, a zebrafish homolog of Chx10, marks ipsilateral descending excitatory interneurons that participate in the regulation of spinal locomotor circuits, J. Neurosci. 26 (2006) 5684–5697. [106] P. Blader, C. Plessy, U. Strahle, Multiple regulatory elements with spatially and temporally distinct activities control neurogenin1 expression in primary neurons of the zebrafish embryo, Mech. Dev. 120 (2003) 211–218. [107] A.N. Ng, T.A. de Jong-Curtain, D.J. Mawdsley, S.J. White, J. Shin, B. Appel, P.D. Dong, D.Y. Stainier, J.K. Heath, Formation of the digestive system in zebrafish: III. Intestinal epithelium morphogenesis, Dev. Biol. 286 (2005) 114–135. [108] J. Shin, H.C. Park, J.M. Topczewska, D.J. Mawdsley, B. Appel, Neural cell fate analysis in zebrafish using olig2 BAC transgenics, Methods Cell Sci. 25 (2003) 7–14. [109] B.B. Kirby, N. Takada, A.J. Latimer, J. Shin, T.J. Carney, R.N. Kelsh, B. Appel, In vivo time-lapse imaging shows dynamic oligodendrocyte progenitor behavior during zebrafish development, Nat. Neurosci. 9 (2006) 1506–1511. [110] T.A. Bayer, J.A. Campos-Ortega, A transgene containing lacZ is expressed in primary sensory neurons in zebrafish, Development (Cambridge, England) 115 (1992) 421–426. [111] M.E. Halpern, J. Rhee, M.G. Goll, C.M. Akitake, M. Parsons, S.D. Leach, Gal4/ UAS transgenic tools and their application to zebrafish, Zebrafish 5 (2008) 97–110. [112] J. Dong, G.W. Stuart, Transgene manipulation in zebrafish by using recombinases, Methods Cell Biol. 77 (2004) 363–379. [113] E. Provost, J. Rhee, S.D. Leach, Viral 2A peptides allow expression of multiple proteins from a single ORF in transgenic zebrafish embryos, Genesis 45 (2007) 625–629. [114] C.B. Moens, T.M. Donn, E.R. Wolf-Saxon, T.P. Ma, Reverse genetics in zebrafish by TILLING, Brief. Funct. Genomic Proteomic 7 (2008) 454–459. E. Kabashi et al. / Biochimica et Biophysica Acta 1812 (2011) 335–345 [115] J.E. Foley, J.R. Yeh, M.L. Maeder, D. Reyon, J.D. Sander, R.T. Peterson, J.K. Joung, Rapid mutation of endogenous zebrafish genes using zinc finger nucleases made by Oligomerized Pool ENgineering (OPEN), PLoS One 4 (2009) e4348. [116] I.G. Woods, A.F. Schier, Targeted mutagenesis in zebrafish, Nat. Biotechnol. 26 (2008) 650–651. [117] X. Ouyang, I.A. Shestopalov, S. Sinha, G. Zheng, C.L. Pitt, W.H. Li, A.J. Olson, J.K. Chen, Versatile synthesis and rational design of caged morpholinos, J. Am. Chem. Soc. 131 (2009) 13255–13269. [118] S. Wu, J.M. Storey, K.B. Storey, Phosphoglycerate kinase 1 expression responds to freezing, anoxia, and dehydration stresses in the freeze tolerant wood frog, Rana sylvatica, J. Exp. Zool. A Ecol. Genet. Physiol. 311 (2009) 57–67. [119] P.A. Morcos, Y. Li, S. Jiang, Vivo-morpholinos: a non-peptide transporter delivers morpholinos into a wide array of mouse tissues, Biotechniques 45 (2008) 613–614 616, 618 passim. [120] C.W. Sipe, M.S. Saha, The use of microarray technology in nonmammalian vertebrate systems, Methods Mol. Biol. 382 (2007) 1–16. [121] A. Hegde, N.C. Qiu, X. Qiu, S.H. Ho, K.Q. Tay, J. George, F.S. Ng, K.R. Govindarajan, Z. Gong, S. Mathavan, Y.J. Jiang, Genomewide expression analysis in zebrafish mind bomb alleles with pancreas defects of different severity identifies putative Notch responsive genes, PLoS One 3 (2008) e1479. [122] L.I. Zon, R.T. Peterson, In vivo drug discovery in the zebrafish, Nat. Rev. Drug Discov. 4 (2005) 35–44. 345 [123] J.R. Yeh, K.M. Munson, K.E. Elagib, A.N. Goldfarb, D.A. Sweetser, R.T. Peterson, Discovering chemical modifiers of oncogene-regulated hematopoietic differentiation, Nat. Chem. Biol. 5 (2009) 236–243. [124] A. Mukhopadhyay, R.T. Peterson, Deciphering arterial identity through gene expression, genetics, and chemical biology, Curr. Opin. Hematol. 15 (2008) 221–227. [125] E. Kabashi, L. Lin, M.L. Tradewell, P.A. Dion, V. Bercier, P. Bourgouin, D. Rochefort, S.B. Hadj, H.D. Durham, C. Vande Velde, G.A. Rouleau, P. Drapeau, Gain and loss of function of ALS-related mutations of TARDBP (TDP-43) cause motor deficits in vivo, Human Mol. Gen. 19 (2010) 671–683. [126] J. Gauthier, N. Champagne, R.G. Lafrenière, L. Xiong, D. Spiegelman, E. Brustein, M. Lapointe, H. Peng, M. Côté, A. Noreau, F.F. Hamdan, A. Piton, A.M. Addington, J.L. Rapoport, L.E. DeLisi, M.-O. Krebs, R. Joober, F. Fathalli, F. Mouaffak, A.P. Haghighi, C. Néri, M.-P. Dubé, M.E. Samuels, C. Marineau, E.A. Stone, P. Awadalla, P.A. Barker, S. Carbonetto, P. Drapeau, G.A. Rouleau, and the S2D team, De novo synaptic scaffolding protein SHANK3 mutations in patients ascertained with schizophrenia, Proc. Natl. Acad. Sci. USA 107 (2010) 7863–7868. [127] J. Tarabeux, N. Champagne, E. Brustein, F.F. Hamdan, J. Gauthier, M. Lapointe, C. Maios, A. Piton, D. Spiegelman, É. Henrion, S2D team, B. Mille, J.L. Rapoport, L.E. DeLisi, R. Joober, F. Fathalli, É. Fombonne, L. Mottron, N. Forget-Dubois, M. Boivin, J.L. Michaud, R.G. Lafrenière, P. Drapeau, M.-O. Krebs, G.A. Rouleau, De Novo Truncating Mutation in KIF17 Associated with Schizophrenia, Biol. Psy. 68 (2010) 649–656. Biochimica et Biophysica Acta 1812 (2011) 346–352 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Zebrafish as a tool in Alzheimer's disease research☆ Morgan Newman a,1, Giuseppe Verdile b,1, Ralph N. Martins b, Michael Lardelli a,⁎ a b Discipline of Genetics, School of Molecular and Biomedical Science, The University of Adelaide, SA 5005, Australia Centre of Excellence for Alzheimer's Disease Research and Care, School of Exercise, Biomedical and Health Sciences, Edith Cowan University, Joondalup, WA 6027, Australia a r t i c l e i n f o Article history: Received 27 November 2009 Accepted 21 September 2010 Available online 12 October 2010 Keywords: Zebrafish Alzheimer's disease Presenilin a b s t r a c t Alzheimer's disease is the most prevalent form of neurodegenerative disease. Despite many years of intensive research our understanding of the molecular events leading to this pathology is far from complete. No effective treatments have been defined and questions surround the validity and utility of existing animal models. The zebrafish (and, in particular, its embryos) is a malleable and accessible model possessing a vertebrate neural structure and genome. Zebrafish genes orthologous to those mutated in human familial Alzheimer's disease have been defined. Work in zebrafish has permitted discovery of unique characteristics of these genes that would have been difficult to observe with other models. In this brief review we give an overview of Alzheimer's disease and transgenic animal models before examining the current contribution of zebrafish to this research area. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Alzheimer's disease Alzheimer's disease (AD) is the most common form of neurodegenerative disease. Globally, the incidence of dementia was estimated in 2001 to be 24.3 million with an estimated increase to 81.1 million in 2040 [1]. Clinically, AD is characterised by progressive memory loss and can include impairment of speech and motor ability, depression, delusions, hallucinations, aggressive behaviour and, ultimately, increasing dependence upon others before death (recently reviewed by Voisin and Vellas [2]). The major neuropathological hallmarks of the disease were first described by Alois Alzheimer as “fibers” and “small military” foci (English translation, [3]) Today these are commonly referred to as neurofibrillary tangles (NFTs) and amyloid plaques. Neurofibrillary tangles are hyperphosphorylated forms of the tau protein, while the major protein component of amyloid plaques is a small peptide referred to as amyloid beta (Aβ). Other forms of lesion have also been found in AD brains (reviewed by Duyckaerts et al. [4]). Alzheimer's disease can be classified broadly into two groups, late onset AD (LOAD, occurring N65 years) and early onset AD (EOAD, occurring b65 years). LOAD accounts for the vast majority of AD cases (95%) and is associated with many risk factors. It is well established that age and possession of the ε4 allele of the APOLIPOPROTEIN gene (APOE ε4) are the major risk factors for most of the population (reviewed by Martins et al. [5]). However, other risk factors such as hormonal changes associated with ageing, diet and lifestyle factors play very important roles. Cardiovascular risk factors such as hypercholesterolemia, increased ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Corresponding author. Tel.: +61 8 83033212; fax: +61 8 83034362. E-mail address: [email protected] (M. Lardelli). 1 Made equal contributions to this paper. 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.09.012 LDL and reduced HDL levels, hypertension, obesity and type II diabetes make important contributions to the risk of developing AD (reviewed by Martins et al. [5]). Genome-wide association studies in humans have identified a number of possible additional loci associated with risk for LOAD (reviewed by Rademakers and Rovelet-Lecrux [6]. In particular, two recent studies identified overlapping sets of candidates (see [7,8] and Table1). Individuals carrying the APOE ε4 allele also bear some risk of developing the early onset form of AD. Overall, EOAD accounts for only a small percentage of AD cases but is the most severe form of the disease. The majority of EOAD cases are familial forms of AD showing dominant inheritance. Familial AD (FAD) is typically associated with an early age of onset as young as 25 years (reviewed by Verdile and Martins [20]). Most of our knowledge of the molecular events underlying the development of Alzheimer's disease comes from FAD since we can use genetic analysis to identify the genes and proteins involved. Mutations at only three loci are known to cause FAD although other loci may yet be found (see review by Rademakers and Rovelet-Lecrux [6]). One locus hosts the Amyloid Precursor Protein (APP) from which Aβ is cleaved by the action of two “secretases,” β-secretase and γ-secretase (see Fig. 1). Mutations in APP alter these cleavages so that Aβ production is increased or more aggregation-prone forms of Aβ are formed (reviewed by Verdile et al., [20]). The remaining two loci host the presenilin genes PSEN1 and PSEN2 with mutations in these genes accounting for 50–70% of EOFAD cases [21]. The presenilin proteins form the catalytic core of protein complexes with γ-secretase activity. Mutations in the presenilins also appear to alter the rate or form of Aβ production and many LOAD associated risk factors also regulate Aβ levels in the brain. Therefore, it is not surprising that the currently dominant hypothesis for the cause of FAD (and LOAD) is the “amyloid hypothesis” that postulates that toxic oligomerisation of Aβ is the initiating factor beginning a cascade of consequences such as synaptic M. Newman et al. / Biochimica et Biophysica Acta 1812 (2011) 346–352 347 directed at neutralising Aβ toxicity (i.e. PBT2) are addressing such issues (reviewed by Wisniewski, 2009 [29]). There are also therapies not directed at Aβ that show promise including the anti-histamine Latrepirdine (formerly known as Dimebon™) which has completed phase IIIa trials and is currently in IIIb trials [30]. The mechanism(s) underlying the benefits of Latrepirdine is unclear but may include improving mitochondrial activity. Another agent that has shown some promise is Rember™ (methylthionium chloride) which reduces tau aggregation (see http:// www.taurx.com/) although this drug was not seen to affect the histological or behavioural phenotypes of transgenic zebrafish expressing mutant human MAPT (see 2009 ICAD conference report at www. alzforum.org/new/detail.asp?id=2203). Methylthionium chloride is better known to those working with zebrafish as methylene blue which is used to inhibit fungal growth and to stain the yellow pteridine-containing pigment cells of fish, xanthophores [31]. Recently, the combination of Dimebon (Latrepirdine) and methylene blue has also been shown to inhibit aggregation of TDP-43 [32], a protein component of ubiquitinated inclusions found in the neurodegenerative diseases amyotrophic lateral sclerosis (ALS) and frontotemporal lobar degeneration (FTLD) (reviewed by Mackenzie [33]). The beneficial outcomes of these non-Aβ directed agents may provide further evidence for alternatives to the amyloid hypothesis, However, it is conceivable that these agents may be acting to ameliorate the consequences of Aβ toxicity e.g. by protecting neurons from insults such as aggregation of Aβ or other proteins. Fig. 1. The γ-secretase complex consists of four protein components including PSEN1 or PSEN2 that reside within lipid bilayers and cleave type 1 transmembrane domain proteins at a catalytic site dependent upon two aspartate residues (stars). APP is initially cleaved either by α- or β-secretase to produce APPα or APPβ. Only β-secretase cleavage permits Aβ peptide production after secondary cleavage by γ-secretase that also releases APP's intracellular domain that can enter the nuclear to contribute to gene regulation. dysfunction, inflammation and oxidative damage causing neuronal dysfunction, degeneration and death (see review in [22]). The amyloid hypothesis is an attractively parsimonious explanation for AD pathogenesis and finds support from numerous phenomena observed in EOFAD and LOAD brains. However, it fails to explain some observations in animal models and the outcomes of a number of human clinical trials of drugs intended to ameliorate Aβ toxicity. For example, the majority of transgenic mice harbouring FAD-causing mutations leading to accumulation of human Aβ in their brains do not show dramatic neurodegeneration [23] (see the discussion in the 2008 paper by Chen et al. [24]). However, these transgenic mice do show cognitive deficits in behavioural tests and show reduction in long-term potentiation, most likely indicating loss in synaptic activity [25]. To obtain significant neurodegeneration or neuronal loss in transgenic mice requires expression of three or more mutations in AD-associated genes. Examples of mouse AD models used currently include “3xTg-AD” mice bearing mutant alleles of PS1, APP and MAPT [26] and “5xFAD” mice bearing three APP and two PS1 mutant alleles [27]. However, the coexistence of three or more mutations in AD-associated genes is not a situation that occurs in humans. A more physiologically relevant animal model is required that more closely mirrors human AD neuropathology including the observed neurodegeneration and neuronal loss (see the critical review by Buxbaum [28]). A number of therapeutic agents aimed at reducing Aβ production or accumulation/aggregation have failed human clinical trials. These include peptide mimetics that inhibit Aβ aggregation (i.e., Alzhemed™,) or agents that either inhibit γ-secretase activity (e.g. LY-411,575 and LY-450,139 ) or target the production of the more neurotoxic Aβ42 (e.g. the NSAID, Flurizan™ and refer to www.alzforum.org regarding information on AD drugs in clinical trials). This has raised questions over the validity of the amyloid hypothesis. However, issues surrounding penetration through the blood–brain barrier and the advanced state of neurodegeneration when these drugs were given may explain their failure. Apparently successful therapeutics currently in phase II/III trials such as the immunotherapeutics Bapinezuimab™ and Solabezumab™ or agents 2. Presenilin activity is involved in many aspects of AD pathology The presenilins proteins are known to interact with both the APP and tau proteins [34–38] providing one possible link between Aβ production and changes in tau phosphorylation. Mice lacking expression of both Psen1 and Psen2 in forebrain tissue show a number of phenotypic effects resembling Alzheimer's disease, i.e. neurodegeneration such as hyperphosphorylation of tau, formation of intracellular tau inclusions, gliosis, expansion of brain ventricles and neuronal death all in the absence of any Aβ production [24]. Some characteristics of Alzheimer's disease, such as increased formation of glia, may result from changes in presenilin activity since γ-secretase complexes cleave not only APP but also a large collection of other transmembrane proteins such as NOTCH, E-cadherin, neuregulin-1 and syndecan (reviewed in Ref. [39]). Presenilins also function in a number of cellular processes, at least some of which appear independent of γ-secretase activity. For example, Kang et al. [40] showed that mouse Psen1 represents a second pathway for phosphorylation of β-catenin independent of Axin. Presenilins also appear to act as Ca2+ leak channels in the endoplasmic reticulum [41]. Presenilins are also found in interphase kinetochores [42] (although this observation has not been replicated) and changes in presenilin activity can cause aneuploidy ([43,44]) and dysregulation of cell cycle proteins [12,45]. Indeed, the activity in γ-secretase of at least some of the proteins comprising these complexes does not appear to be their original function. Homologues of presenilins exist in plants and this kingdom does not display γ-secretase activity [46]. The great majority of the mutations causing FAD are found in PSEN1. Currently over 180 different missense mutations are known (data on mutations are collated at Alzheimer Disease and Frontotemporal Dementia Mutation Database, http://www.molgen.ua.ac.be/ADMutations/). There is also evidence that oligomeric forms of Aβ can feed back onto the activity of γ-secretase ([47]). Thus it is possible that mutations that cause production of more Aβ, or more aggregation-prone forms of Aβ, may affect presenilin function and that changes in presenilin function may then affect tau phosphorylation, cell differentiation (e.g. via Notch signalling), calcium homeostasis, etc. 3. Using zebrafish to study genes mutated in FAD Mice are the dominant model used for modelling of Alzheimer's disease pathology (reviewed in Ref. [48]) but non-mammalian organisms have also 348 M. Newman et al. / Biochimica et Biophysica Acta 1812 (2011) 346–352 enlightened us. Mutation screening in the nematode, Caenorhabditis elegans first showed the connection between the presenilins and Notch signalling [49,50]. Expression of human Aβ in C. elegans [51] and Drosophila [52] has been seen to cause amyloid deposits in muscle and neurodegeneration respectively. Aβ toxicity can also be studied using yeast [53]. Use of these organisms has opened up possibilities for screening drug libraries to find compounds blocking Aβ toxicity. As a model organism, zebrafish presents many advantages for the study of human disease. The vertebrate zebrafish genome and anatomy show only ~420 million years (Myr) of divergence from the human lineage [54] rather than the ~600 Myr of the ecdysozoans (Drosophila and C. elegans, [55]). Therefore, most human genes have clearly identifiable orthologues in zebrafish (however, a genome duplication early in the teleost (bony fish) lineage means that, in many cases, two zebrafish genes together perform the function of an orthologue). The numerous and externally fertilized embryos of zebrafish are easily amenable to methods for manipulating gene and protein activity such as injection of antisense oligonucleotides, mRNAs or transgenes and they can be arrayed in microtitre plates for screening of drug libraries. Cell labelling and transplantation for tracking the behaviour and interactions of living cells is also routine. A good indication of the potential utility of zebrafish for Alzheimer's disease research is the attempts to obtain patent coverage of wide areas of the field (e.g. WO/2006/081539). Zebrafish possess genes orthologous to all the genes known to be involved in Alzheimer's disease (Table 1) The genes appa and appb [9] are duplicates of an ancestral teleost orthologue of human APP. The genes psen1 [11] and psen2 [16] are orthologues of human PSEN1 and PSEN2, respectively. Our laboratory has also identified duplicates of an ancestral teleost orthologue of the MAPT gene encoding tau protein, mapta and maptb [56] (discussion of tau activities in zebrafish including the MAPT-based model of neurodegeneration of Paquet et al. 2009 [57] is to be found in an accompanying paper by Ed Burton). The body of research literature examining the endogenous zebrafish appa, appb, psen1 and psen2 genes is still quite limited. Musa et al. [9] detected widespread and overlapping transcription of both appa and appb by whole mount in situ transcript hybridisation (WISH) from midgastrulation. Later in development these two genes show differing patterns of transcription but, notably, at 24 hpf both are expressed in the developing forebrain (telencephalon) and ventral diencephalon. In addition, appb is expressed in the ventral mesencephalon, a series of nuclei in the hindbrain and in the spinal cord. Joshi et al. [10] recently examined the effects of reduction of appb function through use of antisense morpholino oligonucleotides blocking translation (while they also attempted to reduce appa mRNA translation it is not clear that they were able to do this. Also, their appb analysis utilized only one morpholino so non-specific effects are not excluded). They observed defective convergent extension movements and reduced body length. Interestingly, injection into zebrafish embryos of mRNA encoding normal human APP was more effective at ameliorating the loss of appb translation than an Alzheimer's mutant form (“APPswe” containing the “Swedish” double mutation K595N with M596L [58]). This demonstrates that zebrafish embryos can be exploited for analysis of differences in the activity of different mutant forms of APP [10]. Two papers examining transcriptional control of appb have been published [59,60]. Both emphasise the importance for expression of a regulatory element in the first intron of the gene. Approximately 14 kb of DNA from the promoter region and first intron of the gene can largely recapitulate the embryonic transcription pattern of appb [59]. Attempts to model Aβ toxicity have, so far, proven difficult and no studies have yet been published. The existence of transgenic fish expressing human APPswe in neurons in zebrafish the under the control of the promoter of the elavl3 gene (previously known as HuC) was reported at the 6th European Zebrafish Development and Genetics Meeting in Rome in July 2009 (Hruscha, Teucke, Paquet, Haass and Schmid) but there is no data yet on any toxic effects. In our laboratories we have followed a number of approaches to analysing Aβ toxicity. Simple incubation of zebrafish embryos in media containing Aβ peptide appears to produce effects on neural development as monitored by changes in the patterns of neurons observed in non-transgenic zebrafish. It can also induce cell and embryo death (Frederich-Sleptsova et al. manuscript in preparation). We also attempted to generate a transgenic zebrafish model to facilitate screening for drugs suppressing Aβ toxicity by expressing the 42 amino acid residue form of human Aβ (Aβ42) in the melanophores (melanocytes) that constitute the zebrafish's dark surface stripes. The hope was to create a highly visible but nevertheless viable phenotype in zebrafish larvae that could then be arrayed in microtitre dishes for exposure to various chemical compounds. For this we used a DNA fragment from the promoter of the zebrafish gene mitfa (previously nacre) that can specifically drive transcription in zebrafish melanophores [61]. This was coupled to DNA coding for a secretory signal fused to Aβ42. Unfortunately, fish bearing this transgene only showed a disturbed pigmentation pattern at 16 months of age—too late for use in drug screening [81]. Analysis of presenilin function in zebrafish has tended to focus on psen1 due to the more frequent role played by mutations of its human orthologue, PSEN1 in AD. Zebrafish psen1 sequence and activity were first analysed by Leimer et al. in 1999 [11] who noted extensive conservation of protein primary structure. Gene transcripts are Table 1 Zebrafish genes orthologous to those involved in Alzheimer's disease. Zebrafish orthologue Accession number Zebrafish references appa appb psen1 psen2 NM_131564 NM_152886 NM_131024 NM_131514 [9,10] Clusterin (CLU) Phosphatidylinositol Binding Clathrin Assembly Protein (PICALM) Complement Component (3b/4b) Receptor 1 (CR1) apoeb mapta maptb Clu picalm Unknown NM_131098 N/A N/A NM_200802 NM_200927 Loci for γ-secretase complex members Nicastrin (NCT) Anterior Pharynx Defective 1 Homolog A (APH1a) Anterior Pharynx Defective 1 Homolog B (APH1b) Presenilin Enhancer 2 (PSENEN) ncstn unknown aph1b psenen NM_001009556 [17] NM_200115 NM_205576 [18] [18,19] Loci for dominant mutations in FAD Amyloid Precursor Protein (APP) Presenilin 1 (PSEN1) Presenilin 2 (PSEN2) Loci associated with risk for ADa Apolipoprotein E (APOE) Microtubule-Associated Protein Tau (MAPT) a A large number of additional associated genes have been reported and are reviewed in Rademakers and Rovelet-Lecrux, 2009. [11–15] [12,15,16] M. Newman et al. / Biochimica et Biophysica Acta 1812 (2011) 346–352 apparently present in all cells at all developmental times examined. When zebrafish Psen1 protein expression was driven at high levels in cultured human HEK293 cells the zebrafish protein (as expected) displaced human PSEN1 from γ-secretase complexes indicating sufficient structural conservation to interact with other complex components. However, the C-terminal fragment (CTF) of zebrafish Psen1 generated by the endoproteolysis of the holoprotein (required to activate γ-secretase activity [62]) is sufficiently different in size that it can be distinguished from the CTF of human PSEN1 on western blots. Also, zebrafish Psen1 possesses sufficient primary structural differences from human PSEN1 that its γ-secretase cleavage of APP was abnormal leading to increased production of Aβ42 relative to production of the less aggregation-prone 40 amino acid residue form of Aβ (Aβ40). These authors also demonstrated that mutation of one of the two critical catalytic aspartate residues in zebrafish Psen1 could abolish its γsecretase activity. As expected from the expression of the psen1 transcript, we showed that Psen1 protein can be observed from the earliest stages of development [14]. Interestingly, this did not appear to be the case for zebrafish psen2 that we identified in 2002 [16]. We detected the ubiquitous presence of psen2 transcripts at all stages but an antibody raised against an oligopeptide corresponding to N-terminal fragment (NTF) sequence from Psen2 protein only revealed higher expression of Psen2 after 6 hpf [16]. However, there is evidence from humans [63,64] and from our own unpublished work of alternative splicing and phosphorylation of PSEN2 orthologous proteins so we presume that species of zebrafish Psen2 protein do, in fact, exist at the earliest developmental stages in this organism. The essential role of presenilin activity in Notch signalling during development and the resorption of dead embryos that occurs in mice has hindered the analysis of the roles of presenilin during vertebrate development. Resorption is not a problem in zebrafish so we have been keen to exploit morpholino antisense technology to block the activity of the zebrafish psen1 and psen2 genes. In mice, loss of Psen1 activity is lethal during development and Psen1−/− embryos show a similar phenotype to those lacking mouse Notch1 activity [65,66]. In contrast, mouse Psen2−/− embryos are viable with no obvious developmental defects [67,68] (although adults do display haemorrhaging and histological changes in lung tissue [69]. However, mice lacking both Psen1 and Psen2 activity have a more severe phenotype and loss of γsecretase activity than when only Psen1 activity is absent [67]. When we blocked psen1 translation in zebrafish we observed similarly irregular delineation of somites to that seen in Psen1−/− mouse embryos and the cyclic expression of the her1 gene that controls somite formation was disturbed in the presomitic mesoderm [13]. Interestingly, blockage of translation of zebrafish psen2 transcripts also produces significant changes in embryo development including expansion of brain ventricles at 48 hpf and decreased trunk neural crest formation with increased numbers of neurons at 24 hpf. The latter two defects at least can be attributed to decreased Notch signalling indicating that Psen2 in zebrafish plays a more prominent role in Notch signalling than its mammalian orthologue [15]. Indeed, in unpublished work, van Tijn et al. [70] have shown that zebrafish homozygous for null mutations in psen1 are, nevertheless, viable and fertile indicating that psen2 activity can satisfy all essential psen1 functions, at least in a laboratory setting. The viable homozygous null psen1 mutant zebrafish of van Tijn et al. [70] show decreased cell proliferation and de novo neurogenesis. Such fish also present an opportunity for in vivo analysis of the effects on psen1 function of mutations causing Alzheimer's disease [70,71]. Since presenilin proteins are ubiquitously expressed, simple ubiquitous transgenic expression of mutated psen1 genes in fish lacking endogenous psen1 activity can allow analysis of their function in vivo. The ability to analyse the function of genes involved in FAD in vivo is important since gene regulatory and functional data from cultured cells and in vitro assays is sometimes not consistent with what occurs in the whole organism. This may be due to genetic and epigenetic changes that 349 occur in cultured cells during successive passages and/or immortalisation. Also, cultured cells exist in a highly abnormal environment without the normal three-dimensional context of cellular contact and signalling. This may explain, to some extent, differences that we have observed while analysing the function of presenilins in zebrafish embryos compared to what has been observed from mammalian cell culture. For example, when we overexpressed Psen2 protein in zebrafish embryos at relatively high levels by mRNA injection we did not see a decrease in Psen1 levels [13] whereas transfection of constructs expressing presenilin proteins into cultured mammalian cells is known to reduce the levels of endogenous presenilin proteins [72]. A fertilized zebrafish egg represents a macroscopic single cell in which the activity of one or multiple endogenous genes/proteins can be subtly manipulated to obtain a phenotypic or biochemical readout. We attempted to model the molecular effects of mutations that alter human PSEN1 splicing to cause FAD by altering the splicing of zebrafish psen1 using injected morpholino antisense oligonucleotides. Mutations causing increased levels of a PSEN1 transcript isoform lacking exon 8 or that cause loss of exon 9 can cause a form of FAD with a peculiar histology of plaques lacking neuritic cores and showing a “cotton wool” morphology [73,74]. Unfortunately, when morpholinos binding over the splice acceptor sites of zebrafish exons cognate to human exon 8 or exon 9 were injected we did not observe the desired exon loss from transcripts. Instead we saw inclusion of cognate introns 7 or 8, respectively, in a proportion of transcripts causing truncation of the open reading frames after exons 7 or 8. Interestingly this produced very strong dominant negative effects in the case of truncation after exon 7 sequence but not for truncation after exon 8 (which is very similar to a normal Psen1 NTF). The dominant negative effects included crossinhibition of psen2 activity [14]. Injection of mRNAs encoding Psen1 proteins truncated after exon 7 or exon 8 sequences produced similar results ([14] and unpublished data). Tests with further morpholinos and mRNAs have shown that transcripts producing dominant negative truncated proteins are those that code for part or all of sequence cognate to human PSEN1 exons 6 and 7 but not exon 8. (Curiously one of two nonsense mutations in zebrafish psen1 currently under analysis in the laboratory of Dana Jongejan-Zivkovic falls in this region but there appears to be no dominant negative effect probably due to nonsense mediated decay of the transcript before it can be translated into a truncated protein. (Jongejan-Zivkovic personal communication). The discovery of potently dominant negative forms of truncated presenilin may be important since it has been suggested that a unifying feature of all the 180+ mutations in human PSEN1 may be that they are hypomorphic with respect to γ-secretase activity [75]. It is possible that the truncated forms of numerous proteins that are seen to accumulate in the brains of LOAD individuals may represent a failure of RNA quality surveillance mechanisms such as nonsense mediated decay [76]. This, combined with increased rates of aberrant splicing, etc. seen in aged cells [77] may allow accumulation of dominant negative truncated forms of presenilin in ageing brains that could suppress γsecretase activity and thus contribute to LOAD. A zebrafish model of FAD or LOAD would be very useful for screening for drugs to prevent or ameliorate these diseases. However, even embryo development in non-transgenic zebrafish can provide assays for the actions of drugs relevant to Alzheimer's disease. The drug N-[N(3,5-difluorophenacetyl)-l-alanyl]-S-phenylglycine t-butyl ester (DAPT, [78]) is now commonly used to inhibit γ-secretase activity in zebrafish [79]. Unfortunately, it is commonly and incorrectly assumed to be blocking only Notch signalling which, while very important, is not the only influential signalling pathway facilitated by γ-secretase. For example, as noted above, Joshi et al. [10] showed that inhibition of appb activity causes aberrant convergence-extension movements although it has not been shown that γ-secretase activity is required for this and treatment with DAPT [79,15] or simultaneous inhibition of translation of both Psen1 and Psen2 by morpholino injection [15] does not produce this effect. We have previously shown that numbers of 350 M. Newman et al. / Biochimica et Biophysica Acta 1812 (2011) 346–352 a particular spinal cord cell type, the Dorsal Longitudinal Ascending (DoLA) interneuron, appear dependent upon Psen2 but not Psen1 activity [14,15]. We used this as an assay for Psen2 activity to show that truncated Psen1 could suppress Psen2 activity [14]. Finally, in a publication from the Xia laboratory, Yang et al. [80] used zebrafish embryos to observe the differential effects on development of drugs inhibiting primarily the APP or NOTCH cleavage activities of γ-secretase. Thus, zebrafish embryos can provide a valuable tool for discovery of additional compounds that show differential inhibitory effects on γsecretase. The Xia laboratory has also used injection of different combinations of morpholino antisense oligonucleotides to analyse the effects of inhibiting translation of additional protein components of γ-secretase complexes such as aph1b (orthologous to human APH1B) and psenen (orthologous to human PSENEN and previously named PEN2). In this way they demonstrated that zebrafish Psenen interacts with NF-kappaB [19] and the p53-dependent apoptosis pathway [18]. 4. Concluding remarks Despite many years of intensive research uncertainty still surrounds the fundamental mechanisms underlying AD pathology. There is considerable doubt about the relevance of existing transgenic mouse models of the disease. Our lack of understanding has slowed the development of therapies to treat or prevent AD. The vertebrate genome and neurobiology of zebrafish and our rapidly expanding ability to manipulate gene activity in this organism offer alternative pathways in which to probe the mysteries of AD pathology. In particular, zebrafish are well suited to biochemically-based approaches whereby gene activities are manipulated transiently and then the consequences examined in an otherwise normal cellular environment. As genomewide association studies reveal additional genes involved with development of AD pathology these biochemically-based approaches can rapidly contribute to our understanding of the relevance of these genes while reducing the chance of being misled by experimental results from tissue culture systems that may not reflect in vivo reality. Acknowledgments This work was supported by a project grant from the National Health and Medical Research Council of Australia (453622), the Cancer Council of South Australia (S 24/04) and the School of Molecular and Biomedical Sciences of The University of Adelaide. Professor Martins is supported by grants from the McCusker Foundation for Alzheimer's Disease Research, Department of Veterans Affairs, National Health and Medical Research Council (NHMRC), Centre of Excellence for Alzheimer's disease Research and Care (RNM). Dr Verdile is generously supported by grants from Mr. Warren Milner (Milner English College, Perth, Western Australia) and Ms Helen Sewell. The sources of funding listed above played no role in study design; in the collection, analysis and interpretation of data; in the writing of the report; or in the decision to submit the paper for publication. We are grateful to Prof. Christian Haass and others for valuable comments on the manuscript. The authors declare that they have no conflict of interest in this work. References [1] C.P. Ferri, M. Prince, C. Brayne, H. Brodaty, L. Fratiglioni, M. Ganguli, K. Hall, K. Hasegawa, H. Hendrie, Y. Huang, A. Jorm, C. Mathers, P.R. Menezes, E. Rimmer, M. Scazufca, Global prevalence of dementia: a Delphi consensus study, Lancet 366 (2005) 2112–2117. [2] T. Voisin, B. Vellas, Diagnosis and treatment of patients with severe Alzheimer's disease, Drugs Aging 26 (2009) 135–144. [3] A. Alzheimer, R.A. Stelzmann, H.N. Schnitzlein, F.R. Murtagh, An English translation of Alzheimer's 1907 paper, “Uber eine eigenartige Erkankung der Hirnrinde”, Clin. Anat. 8 (1995) 429–431. [4] C. Duyckaerts, B. Delatour, M.C. Potier, Classification and basic pathology of Alzheimer disease, Acta Neuropathol. 118 (2009) 5–36. [5] I.J. Martins, E. Hone, J.K. Foster, S.I. Sunram-Lea, A. Gnjec, S.J. Fuller, D. Nolan, S.E. Gandy, R.N. Martins, Apolipoprotein E, cholesterol metabolism, diabetes, and the convergence of risk factors for Alzheimer's disease and cardiovascular disease, Mol. Psychiatry 11 (2006) 721–736. [6] R. Rademakers, A. Rovelet-Lecrux, Recent insights into the molecular genetics of dementia, Trends Neurosci. 32 (2009) 451–461. [7] J.C. Lambert, S. Heath, G. Even, D. Campion, K. Sleegers, M. Hiltunen, O. Combarros, D. Zelenika, M.J. Bullido, B. Tavernier, L. Letenneur, K. Bettens, C. Berr, F. Pasquier, N. Fievet, P. Barberger-Gateau, S. Engelborghs, P. De Deyn, I. Mateo, A. Franck, S. Helisalmi, E. Porcellini, O. Hanon, M.M. de Pancorbo, C. Lendon, C. Dufouil, C. Jaillard, T. Leveillard, V. Alvarez, P. Bosco, M. Mancuso, F. Panza, B. Nacmias, P. Bossu, P. Piccardi, G. Annoni, D. Seripa, D. Galimberti, D. Hannequin, F. Licastro, H. Soininen, K. Ritchie, H. Blanche, J.F. Dartigues, C. Tzourio, I. Gut, C. Van Broeckhoven, A. Alperovitch, M. Lathrop, P. Amouyel, Genome-wide association study identifies variants at CLU and CR1 associated with Alzheimer's disease, Nat. Genet. 41 (2009) 1047–1048. [8] D. Harold, R. Abraham, P. Hollingworth, R. Sims, A. Gerrish, M.L. Hamshere, J.S. Pahwa, V. Moskvina, K. Dowzell, A. Williams, N. Jones, C. Thomas, A. Stretton, A.R. Morgan, S. Lovestone, J. Powell, P. Proitsi, M.K. Lupton, C. Brayne, D.C. Rubinsztein, M. Gill, B. Lawlor, A. Lynch, K. Morgan, K.S. Brown, P.A. Passmore, D. Craig, B. McGuinness, S. Todd, C. Holmes, D. Mann, A.D. Smith, S. Love, P.G. Kehoe, J. Hardy, S. Mead, N. Fox, M. Rossor, J. Collinge, W. Maier, F. Jessen, B. Schurmann, H. van den Bussche, I. Heuser, J. Kornhuber, J. Wiltfang, M. Dichgans, L. Frolich, H. Hampel, M. Hull, D. Rujescu, A.M. Goate, J.S. Kauwe, C. Cruchaga, P. Nowotny, J.C. Morris, K. Mayo, K. Sleegers, K. Bettens, S. Engelborghs, P.P. De Deyn, C. Van Broeckhoven, G. Livingston, N.J. Bass, H. Gurling, A. McQuillin, R. Gwilliam, P. Deloukas, A. Al-Chalabi, C.E. Shaw, M. Tsolaki, A.B. Singleton, R. Guerreiro, T.W. Muhleisen, M.M. Nothen, S. Moebus, K.H. Jockel, N. Klopp, H.E. Wichmann, M.M. Carrasquillo, V.S. Pankratz, S.G. Younkin, P.A. Holmans, M. O'Donovan, M.J. Owen, J. Williams, Genome-wide association study identifies variants at CLU and PICALM associated with Alzheimer's disease, Nat. Genet. 41 (2009) 1088–1093. [9] A. Musa, H. Lehrach, V.A. Russo, Distinct expression patterns of two zebrafish homologues of the human APP gene during embryonic development, Dev. Genes Evol. 211 (2001) 563–567. [10] P. Joshi, J.O. Liang, K. Dimonte, J. Sullivan, S.W. Pimplikar, Amyloid precursor protein is required for convergent-extension movements during Zebrafish development, Dev. Biol. 335 (2009) 1–11. [11] U. Leimer, K. Lun, H. Romig, J. Walter, J. Grunberg, M. Brand, C. Haass, Zebrafish (Danio rerio) presenilin promotes aberrant amyloid beta-peptide production and requires a critical aspartate residue for its function in amyloidogenesis, Biochemistry 38 (1999) 13602–13609. [12] M. Newman, B. Tucker, S. Nornes, A. Ward, M. Lardelli, Altering presenilin gene activity in zebrafish embryos causes changes in expression of genes with potential involvement in Alzheimer's disease pathogenesis, J. Alzheimers Dis. 16 (2009) 133–147. [13] S. Nornes, C. Groth, E. Camp, P. Ey, M. Lardelli, Developmental control of Presenilin1 expression, endoproteolysis, and interaction in zebrafish embryos, Exp. Cell Res. 289 (2003) 124–132. [14] S. Nornes, M. Newman, G. Verdile, S. Wells, C.L. Stoick-Cooper, B. Tucker, I. Frederich-Sleptsova, R. Martins, M. Lardelli, Interference with splicing of Presenilin transcripts has potent dominant negative effects on Presenilin activity, Hum. Mol. Genet. 17 (2008) 402–412. [15] S. Nornes, M. Newman, S. Wells, G. Verdile, R.N. Martins, M. Lardelli, Independent and cooperative action of Psen2 with Psen1 in zebrafish embryos, Exp. Cell Res. 315 (2009) 2791–2801. [16] C. Groth, S. Nornes, R. McCarty, R. Tamme, M. Lardelli, Identification of a second presenilin gene in zebrafish with similarity to the human Alzheimer's disease gene presenilin2, Dev. Genes Evol. 212 (2002) 486–490. [17] S.-W. Chong, A. Amsterdam, H.-W. Ang, H. Yang, N. Hopkins, Y.-J. Jiang, Zebrafish Nicastrin is required for mid- and hindbrain development, Dev. Biol. 331 (2009) 466. [18] W.A. Campbell, H. Yang, H. Zetterberg, S. Baulac, J.A. Sears, T. Liu, S.T. Wong, T.P. Zhong, W. Xia, Zebrafish lacking Alzheimer presenilin enhancer 2 (Pen-2) demonstrate excessive p53-dependent apoptosis and neuronal loss, J. Neurochem. 96 (2006) 1423–1440. [19] H. Zetterberg, W.A. Campbell, H.W. Yang, W. Xia, The cytosolic loop of the gammasecretase component presenilin enhancer 2 protects zebrafish embryos from apoptosis, J. Biol. Chem. 281 (2006) 11933–11939. [20] G. Verdile, R. Martins, Molecular Genetics of Alzheimer's Disease, in: D. Wildenauer (Ed.), Molecular Biology of Neuropsychiatric Disorders, SpringerVerlag, Berlin, 2008, pp. 229–276. [21] N. Ercelen, S. Mercan, Alzheimer's disease and genes, Adv. Mol. Med. 1 (2005) 155–164. [22] G. Verdile, S. Fuller, C.S. Atwood, S.M. Laws, S.E. Gandy, R.N. Martins, The role of beta amyloid in Alzheimer's disease: still a cause of everything or the only one who got caught? Pharmacol. Res. 50 (2004) 397–409. [23] E. McGowan, F. Pickford, J. Kim, L. Onstead, J. Eriksen, C. Yu, L. Skipper, M.P. Murphy, J. Beard, P. Das, K. Jansen, M. Delucia, W.L. Lin, G. Dolios, R. Wang, C.B. Eckman, D.W. Dickson, M. Hutton, J. Hardy, T. Golde, Abeta42 is essential for parenchymal and vascular amyloid deposition in mice, Neuron 47 (2005) 191–199. [24] Q. Chen, A. Nakajima, S.H. Choi, X. Xiong, Y.P. Tang, Loss of presenilin function causes Alzheimer's disease-like neurodegeneration in the mouse, J. Neurosci. Res. 86 (2008) 1615–1625. [25] D.S. Woodruff-Pak, Animal models of Alzheimer's disease: therapeutic implications, J. Alzheimers Dis. 15 (2008) 507–521. M. Newman et al. / Biochimica et Biophysica Acta 1812 (2011) 346–352 [26] S. Oddo, A. Caccamo, J.D. Shepherd, M.P. Murphy, T.E. Golde, R. Kayed, R. Metherate, M.P. Mattson, Y. Akbari, F.M. LaFerla, Triple-transgenic model of Alzheimer's disease with plaques and tangles: intracellular Abeta and synaptic dysfunction, Neuron 39 (2003) 409–421. [27] H. Oakley, S.L. Cole, S. Logan, E. Maus, P. Shao, J. Craft, A. Guillozet-Bongaarts, M. Ohno, J. Disterhoft, L. Van Eldik, R. Berry, R. Vassar, Intraneuronal beta-amyloid aggregates, neurodegeneration, and neuron loss in transgenic mice with five familial Alzheimer's disease mutations: potential factors in amyloid plaque formation, J. Neurosci. 26 (2006) 10129–10140. [28] J.N. Buxbaum, Animal models of human amyloidoses: are transgenic mice worth the time and trouble? FEBS Lett. 583 (2009) 2663–2673. [29] T. Wisniewski, AD vaccines: conclusions and future directions, CNS Neurol. Disord. Drug Targets 8 (2009) 160–166. [30] R.S. Doody, S.I. Gavrilova, M. Sano, R.G. Thomas, P.S. Aisen, S.O. Bachurin, L. Seely, D. Hung, Effect of dimebon on cognition, activities of daily living, behaviour, and global function in patients with mild-to-moderate Alzheimer's disease: a randomised, double-blind, placebo-controlled study, Lancet 372 (2008) 207–215. [31] S. Le Guyader, S. Jesuthasan, Analysis of xanthophore and pterinosome biogenesis in zebrafish using methylene blue and pteridine autofluorescence, Pigment Cell Res. 15 (2002) 27–31. [32] M. Yamashita, T. Nonaka, T. Arai, F. Kametani, V.L. Buchman, N. Ninkina, S.O. Bachurin, H. Akiyama, M. Goedert, M. Hasegawa, Methylene blue and dimebon inhibit aggregation of TDP-43 in cellular models, FEBS Lett. 583 (2009) 2419–2424. [33] I.R. Mackenzie, The neuropathology of FTD associated With ALS, Alzheimer Dis. Assoc. Disord. 21 (2007) S44–S49. [34] G. Pigino, A. Pelsman, H. Mori, J. Busciglio, Presenilin-1 mutations reduce cytoskeletal association, deregulate neurite growth, and potentiate neuronal dystrophy and tau phosphorylation, J. Neurosci. 21 (2001) 834–842. [35] A. Takashima, M. Murayama, O. Murayama, T. Kohno, T. Honda, K. Yasutake, N. Nihonmatsu, M. Mercken, H. Yamaguchi, S. Sugihara, B. Wolozin, Presenilin 1 associates with glycogen synthase kinase-3beta and its substrate tau, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 9637–9641. [36] G. Verdile, R.N. Martins, M. Duthie, E. Holmes, P.H. St George-Hyslop, P.E. Fraser, Inhibiting amyloid precursor protein C-terminal cleavage promotes an interaction with presenilin 1, J. Biol. Chem. 275 (2000) 20794–20798. [37] W. Xia, W.J. Ray, B.L. Ostaszewski, T. Rahmati, W.T. Kimberly, M.S. Wolfe, J. Zhang, A.M. Goate, D.J. Selkoe, Presenilin complexes with the C-terminal fragments of amyloid precursor protein at the sites of amyloid beta-protein generation, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 9299–9304. [38] W. Xia, J. Zhang, R. Perez, E.H. Koo, D.J. Selkoe, Interaction between amyloid precursor protein and presenilins in mammalian cells: implications for the pathogenesis of Alzheimer disease, Proc. Natl. Acad. Sci. U. S. A. 94 (1997) 8208–8213. [39] J.V. McCarthy, C. Twomey, P. Wujek, Presenilin-dependent regulated intramembrane proteolysis and gamma-secretase activity, Cell. Mol. Life Sci. 66 (2009) 1534–1555. [40] D.E. Kang, S. Soriano, X. Xia, C.G. Eberhart, B. De Strooper, H. Zheng, E.H. Koo, Presenilin couples the paired phosphorylation of beta-catenin independent of axin: implications for beta-catenin activation in tumorigenesis, Cell 110 (2002) 751–762. [41] H. Tu, O. Nelson, A. Bezprozvanny, Z. Wang, S.F. Lee, Y.H. Hao, L. Serneels, B. De Strooper, G. Yu, I. Bezprozvanny, Presenilins form ER Ca2+ leak channels, a function disrupted by familial Alzheimer's disease-linked mutations, Cell 126 (2006) 981–993. [42] J. Li, M. Xu, H. Zhou, J. Ma, H. Potter, Alzheimer presenilins in the nuclear membrane, interphase kinetochores, and centrosomes suggest a role in chromosome segregation, Cell 90 (1997) 917–927. [43] D.I. Boeras, A. Granic, J. Padmanabhan, N.C. Crespo, A.M. Rojiani, H. Potter, Alzheimer's presenilin 1 causes chromosome missegregation and aneuploidy, Neurobiol. Aging 29 (2008) 319–328. [44] L.N. Geller, H. Potter, Chromosome missegregation and trisomy 21 mosaicism in Alzheimer's disease, Neurobiol. Dis. 6 (1999) 167–179. [45] B. Malik, A. Currais, A. Andres, C. Towlson, D. Pitsi, A. Nunes, M. Niblock, J. Cooper, T. Hortobagyi, S. Soriano, Loss of neuronal cell cycle control as a mechanism of neurodegeneration in the presenilin-1 Alzheimer's disease brain, Cell Cycle 7 (2008) 637–646. [46] A. Khandelwal, D. Chandu, C.M. Roe, R. Kopan, R.S. Quatrano, Moonlighting activity of presenilin in plants is independent of gamma-secretase and evolutionarily conserved, Proc. Natl. Acad. Sci. U. S. A. 104 (2007) 13337–13342. [47] A. Sotthibundhu, A.M. Sykes, B. Fox, C.K. Underwood, W. Thangnipon, E.J. Coulson, Beta-amyloid(1–42) induces neuronal death through the p75 neurotrophin receptor, J. Neurosci. 28 (2008) 3941–3946. [48] M. Newman, I.F. Musgrave, M. Lardelli, Alzheimer disease: amyloidogenesis, the presenilins and animal models, Biochim. Biophys. Acta 1772 (2007) 285–297. [49] R. Baumeister, U. Leimer, I. Zweckbronner, C. Jakubek, J. Grunberg, C. Haass, Human presenilin-1, but not familial Alzheimer's disease (FAD) mutants, facilitate Caenorhabditis elegans Notch signalling independently of proteolytic processing, Genes Funct. 1 (1997) 149–159. [50] X. Li, I. Greenwald, HOP-1, a Caenorhabditis elegans presenilin, appears to be functionally redundant with SEL-12 presenilin and to facilitate LIN-12 and GLP-1 signaling, Proc. Natl. Acad. Sci. U. S. A. 94 (1997) 12204–12209. [51] C.D. Link, Expression of human beta-amyloid peptide in transgenic Caenorhabditis elegans, Proc. Natl. Acad. Sci. U. S. A. 92 (1995) 9368–9372. [52] D.C. Crowther, K.J. Kinghorn, E. Miranda, R. Page, J.A. Curry, F.A. Duthie, D.C. Gubb, D.A. Lomas, Intraneuronal Abeta, non-amyloid aggregates and neurodegeneration in a Drosophila model of Alzheimer's disease, Neuroscience 132 (2005) 123–135. [53] P. Bharadwaj, L. Waddington, J. Varghese, I.G. Macreadie, A new method to measure cellular toxicity of non-fibrillar and fibrillar Alzheimer's Abeta using yeast, J. Alzheimers Dis. 13 (2008) 147–150. 351 [54] V. Ravi, B. Venkatesh, Rapidly evolving fish genomes and teleost diversity, Curr. Opin. Genet Dev. 18 (2008) 544–550. [55] K.J. Peterson, J.B. Lyons, K.S. Nowak, C.M. Takacs, M.J. Wargo, M.A. McPeek, Estimating metazoan divergence times with a molecular clock, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 6536–6541. [56] M. Chen, R.N. Martins, M. Lardelli, Complex splicing and neural expression of duplicated tau genes in zebrafish embryos, J. Alzheimers Dis. 18 (2009) 305–317. [57] D. Paquet, R. Bhat, A. Sydow, E.M. Mandelkow, S. Berg, S. Hellberg, J. Falting, M. Distel, R.W. Koster, B. Schmid, C. Haass, A zebrafish model of tauopathy allows in vivo imaging of neuronal cell death and drug evaluation, J. Clin. Invest. 119 (2009) 1382–1395. [58] M. Citron, T. Oltersdorf, C. Haass, L. McConlogue, A.Y. Hung, P. Seubert, C. VigoPelfrey, I. Lieberburg, D.J. Selkoe, Mutation of the beta-amyloid precursor protein in familial Alzheimer's disease increases beta-protein production, Nature 360 (1992) 672–674. [59] J.A. Lee, G.J. Cole, Generation of transgenic zebrafish expressing green fluorescent protein under control of zebrafish amyloid precursor protein gene regulatory elements, Zebrafish 4 (2007) 277–286. [60] L.A. Shakes, T.L. Malcolm, K.L. Allen, S. De, K.R. Harewood, P.K. Chatterjee, Context dependent function of APPb enhancer identified using enhancer trap-containing BACs as transgenes in zebrafish, Nucleic Acids Res. 36 (2008) 6237–6248. [61] R.I. Dorsky, D.W. Raible, R.T. Moon, Direct regulation of nacre, a zebrafish MITF homolog required for pigment cell formation, by the Wnt pathway, Genes Dev. 14 (2000) 158–162. [62] T. Ratovitski, H.H. Slunt, G. Thinakaran, D.L. Price, S.S. Sisodia, D.R. Borchelt, Endoproteolytic processing and stabilization of wild-type and mutant presenilin, J. Biol. Chem. 272 (1997) 24536–24541. [63] G. Prihar, R.A. Fuldner, J. Perez-Tur, S. Lincoln, K. Duff, R. Crook, J. Hardy, C.A. Philips, C. Venter, C. Talbot, R.F. Clark, A. Goate, J. Li, H. Potter, E. Karran, G.W. Roberts, M. Hutton, M.D. Adams, Structure and alternative splicing of the presenilin-2 gene, Neuroreport 7 (1996) 1680–1684. [64] J. Walter, A. Capell, J. Grunberg, B. Pesold, A. Schindzielorz, R. Prior, M.B. Podlisny, P. Fraser, P.S. Hyslop, D.J. Selkoe, C. Haass, The Alzheimer's disease-associated presenilins are differentially phosphorylated proteins located predominantly within the endoplasmic reticulum, Mol. Med. 2 (1996) 673–691. [65] J. Shen, R.T. Bronson, D.F. Chen, W. Xia, D.J. Selkoe, S. Tonegawa, Skeletal and CNS defects in Presenilin-1-deficient mice, Cell 89 (1997) 629–639. [66] P.C. Wong, H. Zheng, H. Chen, M.W. Becher, D.J. Sirinathsinghji, M.E. Trumbauer, H.Y. Chen, D.L. Price, L.H. Van der Ploeg, S.S. Sisodia, Presenilin 1 is required for Notch1 and DII1 expression in the paraxial mesoderm, Nature 387 (1997) 288–292. [67] D.B. Donoviel, A.K. Hadjantonakis, M. Ikeda, H. Zheng, P.S. Hyslop, A. Bernstein, Mice lacking both presenilin genes exhibit early embryonic patterning defects, Genes Dev. 13 (1999) 2801–2810. [68] H. Steiner, K. Duff, A. Capell, H. Romig, M.G. Grim, S. Lincoln, J. Hardy, X. Yu, M. Picciano, K. Fechteler, M. Citron, R. Kopan, B. Pesold, S. Keck, M. Baader, T. Tomita, T. Iwatsubo, R. Baumeister, C. Haass, A loss of function mutation of presenilin-2 interferes with amyloid beta-peptide production and notch signaling, J. Biol. Chem. 274 (1999) 28669–28673. [69] A. Herreman, D. Hartmann, W. Annaert, P. Saftig, K. Craessaerts, L. Serneels, L. Umans, V. Schrijvers, F. Checler, H. Vanderstichele, V. Baekelandt, R. Dressel, P. Cupers, D. Huylebroeck, A. Zwijsen, F. Van Leuven, B. De Strooper, Presenilin 2 deficiency causes a mild pulmonary phenotype and no changes in amyloid precursor protein processing but enhances the embryonic lethal phenotype of presenilin 1 deficiency, Proc. Natl. Acad. Sci. U. S. A. 96 (1999) 11872–11877. [70] P. van Tijn, J.T. Paridaen, C. van Rooijen, D. Zivkovic, Zebrafish models for familial Alzheimer's disease, 6th European Zebrafish Genetics and Development Meeting, 2009, Rome. [71] D. Zivkovic, Investigator profile: an interview with Danica Zivkovic, Ph.D, Zebrafish 4 (2007) 1–8. [72] G. Thinakaran, C.L. Harris, T. Ratovitski, F. Davenport, H.H. Slunt, D.L. Price, D.R. Borchelt, S.S. Sisodia, Evidence that levels of presenilins (PS1 and PS2) are coordinately regulated by competition for limiting cellular factors, J. Biol. Chem. 272 (1997) 28415–28422. [73] D.M. Mann, A. Takeuchi, S. Sato, N.J. Cairns, P.L. Lantos, M.N. Rossor, M. Haltia, H. Kalimo, T. Iwatsubo, Cases of Alzheimer's disease due to deletion of exon 9 of the presenilin-1 gene show an unusual but characteristic beta-amyloid pathology known as “cotton wool” plaques, Neuropathol. Appl. Neurobiol. 27 (2001) 189–196. [74] J.B. Kwok, G.M. Halliday, W.S. Brooks, G. Dolios, H. Laudon, O. Murayama, M. Hallupp, R.F. Badenhop, J. Vickers, R. Wang, J. Naslund, A. Takashima, S.E. Gandy, P.R. Schofield, Presenilin-1 mutation L271V results in altered exon 8 splicing and Alzheimer's disease with non-cored plaques and no neuritic dystrophy, J. Biol. Chem. 278 (2003) 6748–6754. [75] M.S. Wolfe, When loss is gain: reduced presenilin proteolytic function leads to increased Abeta42/Abeta40. Talking Point on the role of presenilin mutations in Alzheimer disease, EMBO Rep. 8 (2007) 136–140. [76] F.W. van Leeuwen, D.P. de Kleijn, H.H. van den Hurk, A. Neubauer, M.A. Sonnemans, J.A. Sluijs, S. Koycu, R.D. Ramdjielal, A. Salehi, G.J. Martens, F.G. Grosveld, J. Peter, H. Burbach, E.M. Hol, Frameshift mutants of beta amyloid precursor protein and ubiquitin-B in Alzheimer's and Down patients, Science 279 (1998) 242–247. [77] E. Meshorer, H. Soreq, Pre-mRNA splicing modulations in senescence, Aging Cell 1 (2002) 10–16. [78] H.F. Dovey, V. John, J.P. Anderson, L.Z. Chen, P. de Saint Andrieu, L.Y. Fang, S.B. Freedman, B. Folmer, E. Goldbach, E.J. Holsztynska, K.L. Hu, K.L. Johnson-Wood, 352 M. Newman et al. / Biochimica et Biophysica Acta 1812 (2011) 346–352 S.L. Kennedy, D. Kholodenko, J.E. Knops, L.H. Latimer, M. Lee, Z. Liao, I.M. Lieberburg, R.N. Motter, L.C. Mutter, J. Nietz, K.P. Quinn, K.L. Sacchi, P.A. Seubert, G.M. Shopp, E.D. Thorsett, J.S. Tung, J. Wu, S. Yang, C.T. Yin, D.B. Schenk, P.C. May, L.D. Altstiel, M.H. Bender, L.N. Boggs, T.C. Britton, J.C. Clemens, D.L. Czilli, D.K. Dieckman-McGinty, J.J. Droste, K.S. Fuson, B.D. Gitter, P.A. Hyslop, E.M. Johnstone, W.Y. Li, S.P. Little, T.E. Mabry, F.D. Miller, J.E. Audia, Functional gamma-secretase inhibitors reduce beta-amyloid peptide levels in brain, J. Neurochem. 76 (2001) 173–181. [79] A. Geling, H. Steiner, M. Willem, L. Bally-Cuif, C. Haass, A gamma-secretase inhibitor blocks Notch signaling in vivo and causes a severe neurogenic phenotype in zebrafish, EMBO Rep. 3 (2002) 688–694. [80] T. Yang, D. Arslanova, Y. Gu, C. Augelli-Szafran, W. Xia, Quantification of gammasecretase modulation differentiates inhibitor compound selectivity between two substrates Notch and amyloid precursor protein, Mol. Brain 1 (2008) 15. [81] M. Newman, L. Wilson, E. Camp, G. Verdile, R.N. Martins, M. Lardelli, A zebrafish melanophore model of amyloidbeta toxicity, Zebrafish 7 (2010) 155–159. Biochimica et Biophysica Acta 1812 (2011) 353–363 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Zebrafish models of Tauopathy☆ Qing Bai a,b, Edward A. Burton a,b,c,d,e,f,⁎ a Pittsburgh Institute for Neurodegenerative Diseases, University of Pittsburgh School of Medicine, USA Department of Neurology, University of Pittsburgh School of Medicine, USA c Department of Microbiology and Molecular Genetics, University of Pittsburgh School of Medicine, USA d Department of Neurology, VA Pittsburgh Healthcare System, USA e Geriatric Research, Education and Clinical Center, VA Pittsburgh Healthcare System, USA f Division of Movement Disorders, University of Pittsburgh Medical Center, USA b a r t i c l e i n f o Article history: Received 24 January 2010 Accepted 8 September 2010 Available online 16 September 2010 Keywords: Tau Tauopathy Progressive supranuclear palsy Zebrafish Alzheimer's disease Transgenic Neurodegeneration a b s t r a c t Tauopathies are a group of incurable neurodegenerative diseases, in which loss of neurons is accompanied by intracellular deposition of fibrillar material composed of hyperphosphorylated forms of the microtubuleassociated protein Tau. A zebrafish model of Tauopathy could complement existing murine models by providing a platform for genetic and chemical screens, in order to identify novel therapeutic targets and compounds with disease-modifying potential. In addition, Tauopathy zebrafish would be useful for hypothesis-driven experiments, especially those exploiting the potential to deploy in vivo imaging modalities. Several considerations, including conservation of specialized neuronal and other cellular populations, and biochemical pathways implicated in disease pathogenesis, suggest that the zebrafish brain is an appropriate setting in which to model these complex disorders. Novel transgenic zebrafish lines expressing wild-type and mutant forms of human Tau in CNS neurons have recently been reported. These studies show evidence that human Tau undergoes disease-relevant changes in zebrafish neurons, including somato-dendritic relocalization, hyperphosphorylation and aggregation. In addition, preliminary evidence suggests that Tau transgene expression can precipitate neuronal dysfunction and death. These initial studies are encouraging that the zebrafish holds considerable promise as a model in which to study Tauopathies. Further studies are necessary to clarify the phenotypes of transgenic lines and to develop assays and models suitable for unbiased high-throughput screening approaches. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. Published by Elsevier B.V. 1. Introduction The microtubule-associated protein Tau (MAP-τ, ‘Tau’) undergoes biochemical alterations, cellular redistribution, and deposition as insoluble intraneuronal fibrils (Fig. 1), in a variety of neurodegenerative conditions that are collectively termed ‘Tauopathies’. Together, these diseases, which include Alzheimer's disease, progressive supranuclear palsy and other conditions (Table 1), are an important cause of morbidity and mortality, with diverse clinical manifestations. No currently available treatments improve the prognosis of any of these relentlessly progressive diseases. Consequently, investigations aimed at determining the underlying pathophysiology of Tauopathies, and isolating novel therapeutic agents that prevent disease progression, are of great importance. In this review, we consider recent developments concerning the possibility that a zebrafish Tauopathy model might be useful for therapeutic target and drug discovery in vivo. After ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Corresponding author. 7015 BST-3 3501 Fifth Avenue, Pittsburgh, PA 15217, USA. Tel.: +1 412 640 8480; fax: +1 412 648 8537. E-mail address: [email protected] (E.A. Burton). 0925-4439/$ – see front matter. Published by Elsevier B.V. doi:10.1016/j.bbadis.2010.09.004 briefly reviewing current knowledge and murine models of Tauopathy, we discuss the possible advantages of a zebrafish model and whether a truly representative model encompassing key biochemical events underlying Tauopathy can be recapitulated in the zebrafish central nervous system. Finally, we review recent publications demonstrating initial proof of concept that Tauopathy zebrafish models recapitulate core features of the human disorders. 2. The microtubule-associated protein Tau Neurons rely on fast axonal transport to shuttle organelles and macromolecules over long distances, allowing their physiological distribution and turnover within axons and dendrites. Microtubules, which provide the tracks along which molecular motors rapidly transport these diverse cargos, are composed of polymerized tubulin monomers. Assembly of tubulin into microtubules is promoted by microtubule-associated proteins, the first of which to be identified was termed ‘Tau’ (τ was used to denote a factor essential for tubule formation) [1,2]. Tau is expressed widely in neurons, where it is enriched in the axonal compartment [3]. The microtubule-binding domain of Tau localizes to the C-terminal half of the protein [4,5] (Fig. 2). The N-terminal, or projection domain, contains a proline-rich 354 Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 isoforms. In adult human brain, six isoforms are expressed, produced by alternative splicing of exons 2, 3 and 10 (exons 4A, 6 and 8 are not expressed in the CNS). Tau isoforms in the CNS contain either three or four copies of a tandem repeat containing tubulin-binding sequences, referred to as 3R- and 4R-Tau respectively (Fig. 2). Exon 10 encodes the second microtubule-binding repeat, such that inclusion of exon 10 in the mRNA results in translation of 4R-Tau, whereas exclusion of exon 10 results in a transcript encoding 3R-Tau [6,7]. Optional inclusion of exon 2, or exons 2 and 3, gives rise to N-terminal insertions of 29 (‘1N’) or 58 (‘2N’) amino acids respectively [8]. The six resulting CNS isoforms are shown in Fig. 2. Additional isoforms generated by incorporation of exon 4A (‘large Tau’) are expressed in the peripheral nervous system [9]. 2.2. Functions of Tau Fig. 1. Neurofibrillary tangles in Alzheimer's disease. A: Neurofibrillary tangles in Alzheimer's disease prefrontal cortex are demonstrated using the Gallyas silver method [127]. NFTs are seen as numerous argyrophilic (black) fibrillar intraneuronal inclusions (arrows). Neuropil threads, axonal abnormalities also caused by accumulations of fibrillar Tau, are also seen (arrowheads). B: Neurofibrillary tangles in Alzheimer's disease hippocampus are demonstrated by immunohistochemistry, using an antibody (AT8) that detects phospho-S202/T205-Tau. NFTs are seen as abundant immunoreactive (red) intraneuronal inclusions (arrows). Accumulations of phospho-Tau are also apparent in abnormal axonal terminals surrounding amyloid plaques (arrowheads). region and multiple potential serine–threonine phosphorylation sites, and is thought to be involved in interactions with other cellular components. Association of Tau with tubulin promotes microtubule formation [1,2], and association of Tau with microtubules is thought to enhance their stability [10]. 4R-Tau binds to microtubules with greater avidity than 3R-Tau isoforms [4,11,12] (this is attributable to sequences in the first inter-repeat segment that are unique to 4R proteins, rather than to the additional copy of the microtubule-binding tandem repeat [13]), suggesting that expression of 4R-Tau may favor microtubule formation and stability. In healthy adult brain tissue, the 4R-/3R-Tau ratio is approximately 1, and disruption of this balance in the human brain may have serious consequences (see below) [14]. However, there are situations in which the 4R-/3R-Tau ratio may be altered physiologically. For example, during embryogenesis, Tau is exclusively expressed in the brain as the 3R/0N isoform [8]. It has been argued that this might promote plasticity of neuronal processes during development by reducing microtubule stability. Recent work has suggested that Tau interacts with motor proteins involved in axonal transport, provoking different functional consequences for distinct types of motor molecules [15]. Kinesin, which transports cargo towards the distal end of the axon, tended to detach from the microtubule on encountering Tau, whereas the proximally-directed motor Dynein showed more directional reversal than detachment. These effects were apparent in the presence of 3R/0N Tau, and suggest that association of Tau with microtubules may modulate axonal transport in addition to promoting microtubule stability. 2.3. Tau phosphorylation 2.1. The MAPT gene Tau is encoded by the MAPT gene, which is located on chromosome 17 and contains 16 exons. Alternative splicing of the primary transcript leads to a family of mRNAs, encoding different protein Tau undergoes post-translational changes, including physiological serine–threonine phosphorylation [16] and O-glycosylation [17], in addition to tyrosine phosphorylation [18], SUMOylation [19] and pathological nitration [20]. Of these changes, phosphorylation has Table 1 Neurodegenerative diseases associated with prominent Tau pathology. Disease Typical clinical presentations Typical Tau pathology Tau filaments Tau species Etiology Alzheimer's disease (AD) Memory disorder, dysphasia, frontal lobe cognitive–behavioral disorder, dementia Falls, rigidity, oculomotor disorder, cognitive–executive disorder Flame-shaped neurofibrillary tangles Paired helical filaments, straight filaments [123] Straight filaments [124] 4R- and 3R-Tau Predominantly 4R-Tau Mostly unkown, few cases caused by mutations in APP, PS1 or PS2 genes Unknown, linked to H1 haplotype at MAPT locus Twisted ribbon filaments [125] Predominantly 4R-Tau Unknown, linked to H1 haplotype at MAPT locus Helical filaments; straight filaments [126] Variable Predominantly 3R-Tau Unknown Variable (see text) Mutations in MAPT gene encoding Tau Progressive supranuclear palsy (PSP) Corticobasal degeneration (CBD) Dystonia, myoclonus, apraxia, cortical sensory loss, dementia Pick's disease (PiD) Frontal lobe cognitive–behavioral disorder, dementia Fronto-temporal dementia and parkinsonism linked to chromosome 17 (FTDP17) Variable Parkinsonian motor disorder and/or frontal lobe dementia Globose neurofibrillary tangles, tufted astrocytes Ballooned neurons, pre-tangles, astrocytic plaques Pick bodies Variable Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 355 Fig. 2. Isoforms of the microtubule-associated protein Tau. The schematic depicts the six Tau isoforms expressed in the adult human brain, labeled to the left of each protein. Positions of major protein domains are shown above the longest isoform. The N-terminal insertions encoded by exons 2 and 3, and the four tandem repeats (R1–4) in the microtubule-binding domains are indicated. Sites of known phosphorylation are shown above the longest isoform. The table to the right of the figure shows the number of amino acids, predicted molecular mass and electrophoretic mobility of each protein isoform. been most extensively studied, since pathological hyperphosphorylation is associated with Tauopathy [21]. Approximately 45 phosphorylation sites have been identified, which cluster in the proline rich C-terminal domains of the protein [22]. Phosphorylation of Tau, in particular at S262 and S356 within the microtubule-binding domain, promotes its detachment from microtubules, suggesting a potential mechanism in regulating microtubule stability and axonal transport [23,24]. In addition to microtubule-binding affinity, phosphorylation may have other effects, such as modulating binding to the chaperone Pin1 [25] and possibly in regulating putative signal transduction functions of Tau through phosphorylation of SH3 domains [26]. Hyperphosphorylation of Tau is a characteristic feature in Tauopathy specimens. No single phosphorylation site is specific for Tauopathy. Although some phospho-epitopes may be more readily detected in pathological specimens compared with non-Tauopathy material, most of the known phosphorylation sites have been demonstrated in normal biopsy material [16]. Hyperphosphorylation is a quantitative rather than qualitative phenomenon, defined by the degree to which Tau is phosphorylated at multiple sites. Phosphorylation is a dynamic process and Tau is rapidly de-phosphorylated post mortem, accounting for the lack of detectable phospho-epitopes in control autopsy specimens. It is unclear whether a functional abnormality of phosphatase activity is present in Tauopathy samples, or whether the failure of de-phosphorylation in Tauopathy autopsy material is attributable to changes in the properties of the hyperphosphorylated molecule as a phosphatase substrate. During normal development, Tau is present in a highly phosphorylated state [27,28], suggesting that the balance of Tau phosphorylation and dephosphorylation can change physiologically, without necessarily provoking neurodegeneration. phosphorylates all of the residues found in Tauopathy specimens. Furthermore, there are examples in vitro of cooperation between kinases; for example, cdk5-mediated phosphorylation of Tau promotes further phosphorylation by GSK3β [37,38]. Consequently, it is thought that the transition from normal to hyperphosphorylated Tau may be a hierarchical or sequential process dependent on phosphorylation by multiple different kinases. Several different protein phosphatases dephosphorylate Tau in vitro, including PP1, PP2A, PP2B and PP2C. Their specificities are overlapping and their roles in vivo are incompletely understood. PP2A has emerged as a strong candidate, since it appears to be the major brain-derived phosphatase activity directed towards Tau in the rat brain [39] and is localized to microtubules (both through an association with Tau and also directly, which may regulate its activity in vitro [40]). 2.5. Tau deposits in sporadic neurodegenerative diseases Tauopathies are characterized pathologically by the presence of fibrillar deposits of hyperphosphorylated Tau. The histological appearance and distribution of Tau deposits differs between diseases, as does the ultrastructural appearance of Tau filaments that compose neurofibrillary tangles, Pick bodies and other inclusions (Table 1). Formation of neurofibrillary tangles (NFTs) likely occurs through oligomeric intermediates, and there is some evidence that these are the primary pathogenic species, rather than mature NFTs (see below). The predominant Tau isoforms deposited in neurons differ between diseases. In Alzheimer's disease, SDS-PAGE analysis showed three major 2.4. Tau kinases Several cellular kinases are able to phosphorylate Tau in vitro or in cell culture (Table 2; reviewed in [22]). Many of these phosphorylate multiple residues of Tau, and individual residues can be phosphorylated by multiple different kinases, suggesting that Tau is a promiscuous kinase substrate. Consequently, it is difficult to be certain which kinase(s) are responsible for physiological or pathological phosphorylation of Tau in vivo. Kinases that have emerged as strong in vivo candidates include glycogen synthase kinase-3β (GSK3β) [29–33] and cyclin-dependent protein kinase-5 (cdk5) [34,35]; both are expressed abundantly in neurons and associate with microtubules (cdk5 is anchored to microtubules by Tau [36]). However, it is likely that more than one kinase participates in the events leading to hyperphosphorylation, since no single kinase Table 2 Phylogenetic conservation of kinases implicated in tau physophorylation. Tau Kinase Protein accession numbers % identity % consensus Human Zebrafish Glycogen synthase kinase 3β (GSK3β) Cyclin-dependent kinase 5 (cdk5) Mitogen activated protein kinase 1 (MAPK1, ERK1) Microtubule Affinity Regulating Kinase 1 (MARK1) Casein kinase 1α (CK1α) NP_001139628 NP_571456 94.5 97.4 AAP35326 AAH85381 96.2 98.6 NP_620407 BAB11813 90.8 93.5 CAH72462 AAI55560 79.3 85.5 NP_001883 NP_694483 99.3 99.3 356 Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 Tau immunoreactive bands of 68, 64 and 60 kDa present in detergentextracted material containing paired helical filaments [41]. After dephosphorylation of the same material, all six Tau isoforms were resolved by electrophoresis, showing that hyperphosphorylation of Tau was responsible for the alteration in its electrophoretic mobility, and that AD filaments are composed of all six Tau isoforms. Similar analyses of specimens derived from other types of Tauopathy have revealed different results. In progressive supranuclear palsy and cortico-basal degeneration, two major Tau bands of 68- and 64 kDa are derived from hyperphosphorylation of 4R-Tau species [42,43]. Conversely, material derived from Pick's disease brain shows two bands of 64- and 60-kDa, dephosphorylation of which reveals the presence of 3R-Tau species [43,44]. Interestingly, genetic data shows association between an extended haplotype of markers across the MAPT locus, and PSP/CBD [45]; recent data shows that this haplotype favors the splicing of exon 10+ transcripts encoding 4R-Tau [46], suggesting that geneticallydetermined relative over-production of 4R-Tau might be one factor in the pathogenesis of these sporadic diseases. 2.6. Hereditary fronto-temporal dementia and Parkinsonism linked to chromosome 17 Definitive evidence that the development of Tau pathology may be mechanistically important in common sporadic Tauopathies comes from the study of rare families, in which mutations within the MAPT gene are sufficient to cause a neurodegenerative disease with prominent Tauopathy [14,47]. FTDP17 is inherited as an autosomal dominant trait and can manifest as a Parkinsonian movement disorder or as a cognitive–behavioral disorder, resembling other forms of fronto-temporal dementia (reviewed in [48]). To date, 39 different pathogenic MAPT mutations have been identified in FTDP17 families. Mutations clustered around the 5′ boundary of intron 10 alter the regulation of splicing, most frequently resulting in relative over-production of exon 10+ transcripts and deposition of hyperphosphorylated 4R-Tau isoforms. Missense mutations within exon 10 also provoke a 4R-Tauopathy, by changing the biophysical properties of 4R-Tau without altering the ratio of splice isoforms; these mutations also result in 4R-Tau deposition. Finally, mutations outside exon 10 are expressed in both 3R- and 4R-Tau, but may provoke selective deposition of 3R- or 4R-, or both isoforms of Tau. A variety of mechanisms has been proposed to account for the pathogenicity of MAPT missense mutants, including promotion of Tau fibril formation, impairment of microtubule-binding and introduction or removal of key phosphorylation sites [49]. Many of the mutants show multiple functional abnormalities, such as alterations in 4R-/3R-splice ratio and reduced propensity to promote the stabilization of microtubules. Although FTDP17 is an uncommon disease, the discovery of pathogenic MAPT mutations has provided important confirmation that primary abnormalities of Tau, including alterations in the relative abundance of 3R- and 4R-isoforms, can result in neurodegeneration and Tauopathy. Given the close similarity between FTDP17 and sporadic Tauopathies in some cases, it is reasonable to conclude that abnormalities of cellular Tau metabolism may be central to neurodegeneration in common sporadic Tauopathies. Consequently, study of model systems in which Tauopathy is induced by molecular manipulations that mimic the genetic mechanisms of FTDP17 might yield important insights into pathophysiology and identify potential treatment targets for common sporadic Tauopathies. 3. Transgenic mouse models of Tauopathy A variety of different transgenic lines, over-expressing wild-type or FTDP17 mutant human Tau, have been described (reviewed in [50,51]). Several important themes that have emerged from studies of these animals are briefly considered here. First, dissociation between NFT formation and neuronal dysfunction or death was demonstrated in some of these models, suggesting that Tau hyper-phosphorylation, relocalization or oligomerization may be more critical for the emergence of phenotypic abnormalities in these models than formation of NFTs. Abnormalities of hippocampal synaptic protein expression and physiology preceded the formation of NFTs in mice expressing P301S mutant Tau [52]. Studies using a different model in which P301L Tau was expressed under an inducible promoter showed that suppression of transgene expression after the onset of phenotypic abnormalities prevented neuronal loss and behavioral abnormalities but did not affect neurofibrillary tangle formation [53], and there was regional dissociation between NFT formation and cell loss [54]. In addition, immunization of Tau/Aβ transgenic mice reduced soluble Tau in the brain but did not affect NFT formation — the animals showed phenotypic improvement, suggesting that neurobehavioral abnormalities were attributable to a soluble Tau species rather than NFT [55]. Second, Tau transgenic mice have provided a powerful means to test the interactions between different biochemical pathways implicated in neurodegeneration. For example, evidence of an in vivo interaction between Tau and Aβ in the pathogenesis of Alzheimer's disease was demonstrated by intracerebral inoculation of the Alzheimer's disease-associated amyloid peptide Aβ1–42, which enhanced the formation of NFTs in a mouse P301L Tau-expressing model [56]. Furthermore, P301L Tau/APP double transgenic animals showed exacerbation of NFT pathology in the presence of the APP transgene, whereas Aβ plaque formation was unaffected by the overexpression of mutant Tau, suggesting that β-amyloid formation or another function of APP might be upstream of Tauopathy in this model [57]. Finally, murine models have provided powerful means to test hypotheses concerning experimental treatment approaches, and biochemical interventions aimed at understanding disease pathogenesis in vivo (reviewed in [58]). For example, administration of Li+, an inhibitor of GSK3, reduced hyperphosphorylation of Tau in vivo and ameliorated some phenotypic abnormalities [31], and another kinase inhibitor targeting cdk5, GSK3 and MAPK1 delayed the onset of a motor deficit in Tau transgenic animals [59]. Other studies have deployed pharmacological interventions in order to clarify the potential therapeutic utility of interventions directed at restoring loss of Tau microtubule-stabilizing function [60], enhancing clearance of pathological Tau species from neurons by inhibiting refolding functions of the HSP90 complex to promote Tau degradation [61], and curtailing a potentially pathogenic neuroinflammatory response by targeting microglial activation using immunosuppressive drugs [52]. Together, these observations show that experimental genetic manipulations, which mimic the mechanisms underlying FTDP17 in humans, provoke phenotypic changes relevant to human Tauopathy in transgenic animals. These models have been valuable for hypothesis-driven experiments aiming to elucidate the pathogenesis of Tauopathy and for testing putative therapeutic approaches. 4. Why would a zebrafish Tauopathy model be useful? The zebrafish is a small freshwater fish that has been used extensively in laboratory studies of vertebrate development. Fish embryos develop externally allowing direct observation of embryogenesis. Many morphological and molecular parallels are shared between fish and other vertebrates allowing insights gained in the zebrafish model to be applied to other systems. Stemming from the use of zebrafish in developmental studies, an extensive array of experimental methodologies has been developed. Straightforward techniques are available for over-expression [62,63] or targeted knockdown of genes of interest [64], allowing the consequences of altered gene function to be determined readily in vivo. The deployment of fluorescent reporters [65] allows direct observation Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 357 of morphology or physiology of individual cells in vivo [66–68]. The ability of zebrafish to breed regularly and produce sizeable clutches of embryos, and the practicable housing of significant numbers of animals, has allowed development of techniques for large-scale genetic [69–72] and chemical modifier [73–75] screens. These have provided numerous novel molecular insights into development, through unbiased phenotype-driven approaches. More recently, it has been demonstrated that intact zebrafish larvae provide an in vivo neurobehavioral platform suitable for screens to find chemical modifiers of vertebrate behavior [76]. It is possible that applying these types of analysis to zebrafish models of neurodegenerative diseases would provide novel molecular insights into disease pathogenesis, and allow identification of potentially therapeutic compounds. In particular, screening studies carried out in the intact vertebrate CNS in vivo might uncover drug targets that are not present in cell culture models commonly used for discovery-driven approaches, such as pathogenic mechanisms that are not cell-autonomous, or which are only expressed in enddifferentiated neurons. Consequently, there has been significant recent interest in the development of zebrafish models of Tauopathy. to functional sub-specialization, suggesting that the zebrafish may be a powerful tool in which to dissect multiple cellular roles of Tau. The two proteins encoded by the zebrafish genes have not yet been detected, but both mapta and maptb mRNAs were expressed in the developing CNS [97]. A complex pattern of alternative splicing of the mapta and maptb transcripts suggests that, like human Tau, zebrafish Tau isoforms with different numbers of microtubule-binding repeats are expressed in the CNS, and larger forms of Tau are expressed in the peripheral nervous system. Interestingly, mapta gives rise to transcripts encoding 4–6 microtubule-binding repeats, whereas maptb is predominantly expressed as a 3-repeat isoform, raising the fascinating possibility that conserved functions of mammalian 3R- and 4R-Tau are distributed between the two zebrafish genes [97]. 5. Is the zebrafish a suitable organism in which to study Tauopathy? 6.1. Techniques for establishment of transgenic lines The proposed use of a small aquatic creature to study complex human neurobehavioral diseases naturally raises questions about relevance — will it be possible to generate a truly representative model, with potential to yield important insights into the human disorder? A variety of considerations suggest that the zebrafish may be a good model system in which to study human neurological diseases. With respect to the human CNS, the zebrafish shows conservation of basic brain organization [77] and many key neuroanatomical [78,79] and neurochemical [80,81] pathways of relevance to human disease; the zebrafish CNS contains microglia [68], cells with astrocytic properties [82,83], oligodendrocytes and myelin [84–87], and a blood–brain barrier [88], all of which have been implicated in the pathogenesis of neurological disease. These observations imply that the tissue environment in the zebrafish CNS may present appropriate conditions in which to recapitulate pathological changes representative of the human disorders. In addition, there is a striking degree of phylogenetic conservation of genes implicated in the pathogenesis of neurodegenerative diseases, and a variety of other CNS functions [89–96]. This suggests that it will be possible to provoke neurodegeneration in zebrafish through conserved biochemical pathways that are sufficiently close to those underlying human disorders that findings in the zebrafish model will yield relevant insights into diseases mechanisms. It is noteworthy that the zebrafish genome contains highly conserved orthologues of each of the kinases implicated in Tau phosphorylation (Table 2 and Fig. 3), which may be of particular relevance to Tauopathy in view of the proposed central role of hyperphosphorylation in pathogenesis. Similar considerations apply to other conserved pathways, and it seems likely that findings in zebrafish models will be representative of the mechanisms underlying the pathogenesis of human Tauopathies and will therefore be relevant to their understanding. Transgenic zebrafish can be generated by micro-injecting linearized plasmid DNA into the cytoplasm of one-cell stage embryos [100], but this gives rise to inefficient genomic integration and significant mosaicism. Two technical advances, meganucleases and transposons, have substantially improved the efficiency by which foreign DNA can be introduced into the zebrafish genome. (i) The meganuclease I-sce1 is an intron-encoded endonuclease from Saccharomyces cerevisiae [101], which cleaves DNA at an 18 bp sequence-specific recognition site not found within the zebrafish genome. In meganuclease-mediated transgenesis, the transgene plasmid is designed so that the expression cassette is flanked by I-sce1 sites, and is injected into zebrafish embryos with I-sce1 enzyme. Integration of a low transgene copy number usually occurs at a single site, with substantially enhanced efficiency, and reduced mosaicism, compared with naked DNA injection. This improves the rate of transmission of the transgene from the F0 to the F1 generation [102] and results in simple Mendelian transmission of transgenes in subsequent crosses [103]. (ii) Type 2 transposons are mobile DNA elements encoding a transposase, which mediates excision of the transposon from the genome and its re-insertion at another location. The Tol2 transposon, found in the genome of medaka [104], has been engineered by deletion of the transposase gene to form a nonautonomous mobile element, which can insert into the genome in the presence of transposase supplied in trans [105]. In this approach to transgenesis, the transgene expression cassette is flanked by transposon elements, and injected into zebrafish embryos with mRNA encoding transposase, resulting in highly efficient, usually multiple, single-copy integration events. The technique has been used increasingly since its introduction and subsequent refinement [62], because it is necessary to inject and screen a relatively small number of embryos in order to identify stable transgenic lines. 6. Tools for the generation and analysis of zebrafish Tau models Generation of transgenic Tau zebrafish requires availability of appropriate methods for introducing exogenous DNA into the germline, and resources to direct MAPT transgene expression in the desired temporal and spatial pattern. 6.2. Promoter resources for generation of neurodegeneration models 5.1. Zebrafish paralogues of MAPT Recent work has identified two paralogues of MAPT in zebrafish, annotated mapta and maptb [97]. Analysis of homology between zebrafish and mammalian MAP-encoding genes, and examination of synteny between zebrafish mapt and mammalian MAPT loci, strongly suggest that the two zebrafish genes arose from duplication of an ancestral mapt locus. A genome duplication event is thought to have taken place during the evolution of teleosts [98]; consequently, there are dual zebrafish paralogues of many mammalian genes [92,99]. It is possible that divergent evolution of the two mapt genes has given rise Most transgenic zebrafish lines have been constructed by expressing a cDNA under control of a characterized heterologous promoter fragment; cis-acting regulatory regions derived from zebrafish genes are usually employed for this purpose, because they may be associated with more reliable transgene expression [106,107]. For the generation of a Tauopathy model, desirable attributes for the promoter would include pan-neuronal activity and sufficient expression to provoke pathology. Three zebrafish promoter elements that drive transgene expression widely in CNS neurons have been described. (i) A neuronal enhancer derived from the gata2 promoter was expressed at high levels in the 358 Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 Fig. 3. Phylogenetic conservation of kinases implicated in Tau phosphorylation. The sequence figure shows the striking degree of amino acid conservation between human and zebrafish (A) GSK3β and (B) cdk5. Each panel shows the Human (upper) and zebrafish (lower) protein sequences, aligned using the ClustalW algorithm. Amino acids are shaded to show identical, conserved and non-similar residues. The catalytic domain of each kinase is indicated by underlining, and important functional sites labeled as indicated. developing CNS [108]. (ii) The zebrafish huc gene encodes an RNA binding protein, HuC/D, commonly used as an early neuronal marker [109] — a 2.8 kb fragment of the proximal 5′ flanking region was sufficient to drive pan-neuronal expression in embryos [110]. (iii) The eno2 gene, encoding the neuron-specific γ-enolase isoenzyme, is expressed at low levels until 60–72 h post-fertilization, after which high level expression persists into adulthood. The eno2 regulatory region is complex, containing an untranslated first exon, and an intronic CpG island that appears important for promoter activity. A 12 kb fragment of the promoter, including the first intron, was active in driving reporter gene expression in neurons, including many populations of relevance to disease, throughout the brain and spinal cord from 36–48 h post-fertilization through adulthood [111], and in the retina and visual pathways [103] (Fig. 4). The optimal promoter for construction of Tauopathy models is unclear, and each of these promoters has been deployed in studies of transgenic Tau zebrafish (see below). It is likely that further development of promoter resources and other approaches to generation of transgenic animals will yield other useful reagents for this application. ductive potential or preventing survival to sexual maturity. One possible solution would be deployment of a conditionally-expressing system, for example the binary Gal4–UAS system previously employed in Drosophila genetics. This system has been successfully used in zebrafish, with recent modifications attempting to address toxicity issues caused by the viral trans-activation domain usually fused to Gal4, and inactivation of the UAS tandem repeats [112,113]. Isolation of useful driver lines for generation of neurodegenerative disease models is ongoing, and the optimal strategy for deployment of this technology is uncertain. Driver lines could be constructed by expressing the Gal4-transactivator fusion protein using small promoter fragments similar to the approach for generating cDNA transgenic animals; this has been used successfully with the huC promoter in a larval Tau model [114]. In addition, the highly efficient genomic integration of transgenes using the Tol2 transposon has allowed the deployment of approaches in which random integration of a gal4-expressing enhancer trap gives rise to robust stable expression of Gal4 under the endogenous regulatory elements of the gene of insertion [115]. Several such lines have been constructed and are currently being further evaluated. 6.3. Special considerations for expression of transgenes that provoke neurodegeneration in larvae 6.4. Assays for fully exploiting transgenic models In order to fully exploit the potential of the zebrafish for highthroughput screening approaches, it may be necessary to drive Tau expression at sufficiently high levels that disease pathogenesis occurs during larval stages of development, when zebrafish can be accommodated in 96-well plates. However, development of neurodegeneration at this early time point could provide strong selection against establishment of transgenic lines, by compromising repro- The power of screening paradigms might be enhanced by deployment of assays that exploit unique properties of the zebrafish model. First, direct visualization of neuronal populations of interest in the zebrafish brain might allow non-invasive automated imaging to ascertain the morphology, integrity or viability of target cells as an assay end point. For example, the Tauopathies PSP and CBD are associated with severe depletion of dopaminergic neurons causing a Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 359 Fig. 4. The zebrafish eno2 promoter. The micrograph shows a single confocal plane through the head region of a Tg(eno2:egfp) zebrafish larva at 5 days post-fertilization, illustrating widespread neuronal expression of the eno2 promoter. Rostral is to left, dorsal up. GFP expression appears green; the section was counterstained using a blue nuclear marker to facilitate identification of anatomical landmarks. Key: Tel, telencephalon; TeO, optic tectum; Ce, cerebellum; MdO, medulla oblongata; SC, spinal cord; L, ocular lens; PRL, photoreceptor layer of retina; IPL, inner plexiform layer of retina; GCL, ganglion cell layer of retina; TG, trigeminal ganglion; OC, otic capsule; PLLG, posterior lateral line ganglion; LLN, axons of lateral line nerve. Parkinsonian movement disorder that characterizes these conditions [116], and it would be of considerable interest to identify compounds that protect dopaminergic neurons from abnormal forms of Tau. Genetic labeling of dopamine neurons using fluorescent reporters is ongoing, and has proved to be complicated. The promoter regions of dopamine-neuron-specific genes are large and complex, such that manageable fragments of the th [117,118] or slc6a3 [119] promoters have not shown dopaminergic neuron-specific activity in vivo. Alternative approaches are being undertaken, including identification of conserved enhancers, using large genomic constructs likely to contain more cis-acting elements, or adopting an enhancer trap approach to identify genomic integrants that recapitulate endogenous gene expression patterns. The latter approach has yielded a single transgenic line in which an integration event in the vmat gene yielded zebrafish with GFP-expressing catecholaminergic neurons [67]. There is a large literature on the development of promoter resources allowing genetic labeling of other specific neuronal populations using fluorescent reporters, which is outside the scope of the present review. Second, neuronal dysfunction presumably precedes cell death in Tauopathy, and cells in earlier stages of pathological evolution might present a more tractable therapeutic target than those close to demise. It would be of considerable value to develop assays that measure neuronal dysfunction, and which are sensitive to changes in disease progression earlier in the evolution of pathology. Since genetic and chemical screening against a zebrafish Tauopathy model would be carried out in vivo, it might be possible to use automated motor behavioral assays as a surrogate marker of neuronal function. Consequently there is much current interest in establishing reproducible and well-characterized behavioral assays for neurodegenerative phenotypes. A recent report showed that an automated motor behavioral test could be used to identify agents with neuropharmacological properties in a chemical screen [76]. Assays of motor function have been previously deployed in order to detect changes in behavior caused by alterations in neurochemistry and dopamine cell integrity secondary to Parkinson's disease-associated toxins [81,120]. It may be possible to use a similar approach to investigate changes in behavior preceding demonstrable cell loss in a zebrafish Tauopathy model, in order to develop an assay to identify chemical modifiers of early pathological progression. 7. Current transgenic zebrafish models of Tauopathy Three recent publications have described the first proof-of-principle experiments showing that expression of human MAPT transgenes can provoke relevant phenotypes in zebrafish. In the first publication, a 4R-Tau-GFP fusion protein was transiently over-expressed in zebrafish larvae, using a gata2 promoter element [121]. In cultured cells, the fusion protein was phosphorylated and associated with the cytoskeleton, similar to native Tau; in vitro studies showed that the fusion protein aggregated, validating its use in constructing an in vivo model of NFT formation. Zebrafish embryos microinjected with the gata2:MAPT-egfp construct, showed mosaic neuronal expression of the fusion protein. Some neurons showed accumulation of fluorescent fibrillar structures resembling neurofibrillary tangles, in the cell body and proximal axon. The human Tau-GFP fusion protein was shown to be phosphorylated in the zebrafish brain by western blot detection of a phospho-S396/S404 epitope, and expression of glycogen synthase kinase-3 in the zebrafish embryo and adult brain was demonstrated by western blot. This initial study validated the use of a GFP fusion protein to monitor evolution of tangle pathology in vivo, and suggested that human Tau is phosphorylated in the larval zebrafish, confirming the prediction that there is sufficient phylogenetic conservation of endogenous zebrafish kinases to modify the human protein. No stable lines were derived in this study. Subsequently, stable transgenic zebrafish were constructed expressing human 4R/0N Tau [111], the most abundant Tau species deposited in progressive supranuclear palsy [43,44]. The transgene was expressed under transcriptional control of the eno2 promoter [111]. The full phenotype of these transgenic fish has not yet been reported, since the initial paper focused on characterization of the novel eno2 promoter. Abundant expression of human 4R/0N Tau was found in the CNS, persisting into adulthood (Fig. 5A). In the adult brain, refractile Tau accumulations resembling neurofibrillary tangles were found within neuronal cell bodies and proximal axons in CNS regions of pathological relevance to PSP, including the optic tectum (Fig. 5B). Stable expression of the Tau transgene into adulthood in this model will allow analyses designed to test whether age-related pathology accumulates in these transgenic lines, similar to the progressive changes seen in the human disease and recapitulated in some murine models, a potentially important step in validating the model system. Furthermore, widespread neuronal expression of the transgene may allow examination of factors involved in cellular vulnerability to Tauopathy. More recently, exploitation of the Gal4–UAS system allowed generation of a Tauopathy model showing a larval phenotype [114]. This is potentially an important advance, since the development of phenotypic abnormalities at larval stages of development may allow high-throughput screening, as discussed above. Human 4R/2N-Tau, harboring a P301L mutation, was expressed from a novel bidirectional UAS promoter, allowing simultaneous expression of a red fluorescent reporter in Tau-expressing cells [114]. The high levels of mutant P301L Tau expression provoked by the huc:gal4-vp16 driver induced a transient motor phenotype during embryogenesis, likely caused by peripheral motor axonal developmental abnormalities. At later time 360 Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 Fig. 5. Human Tau expression in Tg(eno2:MAPT) zebrafish neurons. A: A western blot was made with protein lysate from wild-type zebrafish brain (lane 1), control post-mortem human cortex (lane 2) and Tg(eno2:MAPT) brain (lane 3). The blot was probed using an antibody to human Tau (upper panel) followed by an antibody to β-actin (lower panel). Abundant expression of human 4R-Tau is seen in transgenic zebrafish compared with the six isoforms in normal human brain [111]. B: Sections of Tg(eno2:MAPT)zebrafish brain were labeled using the antibody to human Tau used in the western blot experiment shown in panel A and a histochemical reaction yielding a red product. The micrograph shows a brainstem neuron with dense Tau immunoreactivity in the cell body and proximal axon, resembling neurofibrillary tangles shown in Fig. 1B [111]. points, after recovery of motor function, enhanced cell death was observed in the spinal cord of transgenic animals. Abnormal Tau conformers were detected using an antibody, MC1, which recognizes a discontinuous epitope present in Tauopathy tissue [122], as early as 32 h post-fertilization; by 5 weeks of age, argyophilic material was detected in the spinal cord, but the animals were otherwise phenotypically unremarkable at this time point. Subsequent loss of transgene expression prevented the examination of later pathological changes. Tau phosphorylation was detected as early as 32 h postfertilization. Phospho-Tau specific antibodies labeled cells differentially; some phospho-epitopes, such as phospho-T231/S235 and phospho-S396/S404 were detected widely in Tau-expressing neurons, whereas others such as phospho-S422 and phospho-S202/T205 were initially present in only a small subset of Tau-expressing neurons. Between 2 and 7 days post-fertilization, however, expression of all of these epitopes became widespread in Tau-expressing cells, suggesting that Tau had accumulated multiple phosphorylations in a sequential progression of biochemical changes. Tau phosphorylation was reduced by application of novel high-potency GSK3β inhibitors that were designed to target the human enzyme, confirming that extensive homology between human and zebrafish kinases allows small molecules to interact with orthologous targets from either species. This important study was the first detailed phenotypic description of a stable zebrafish Tauopathy model and showed proof of principle that biochemical changes characteristic of human Tauopathy can be recapitulated in larval zebrafish and modulated by chemical inhibitors. 8. Conclusions — what have we learned and what next? Examination of the first publications allows preliminary assessment of the degree to which transgenic zebrafish models of Tauopathy seem representative of human disorders. Evidence so far suggests that processing of over-expressed human Tau in zebrafish CNS neurons results in similar biochemical and pathological changes to those found in human Tauopathies. These changes include: aberrant localization of Tau to the somato-dendritic compartment; phosphorylation, with an ordered (and possibly sequential) acquisition of phospho-epitopes; adoption of abnormal conformations associated with Tauopathy; and deposition as argyrophilic material. The availability of zebrafish lines that express Tau in the brain through adulthood will allow more detailed biochemical studies of Tau in the zebrafish CNS to characterize solubility, fibril morphology and further clarify phosphorylation events. However, these data are encouraging that relevant pathways are sufficiently phylogenetically conserved between human and zebrafish to allow recapitulation of cellular processing events relating to Tau in the zebrafish model. The pathophysiological consequences of human Tau transgene expression in zebrafish CNS neurons are less clear. This is partly because a detailed phenotype has only been reported for a single transgenic Tau zebrafish line [114]. The motor axonal developmental abnormality, motor phenotype and cell death reported in this model have an uncertain relationship to the neuronal dysfunction and neurodegeneration that characterizes human Tauopathy, on account of the onset and subsequent resolution of abnormalities during early development, and the atypical anatomical distribution in the spinal cord and motor neurons. One recurrent finding from the murine transgenic disease model literature is that phenotypes exhibited by cDNA transgenic animals are critically dependent on the promoter elements used to drive transgene expression. The temporal and spatial expression pattern of the huC promoter used in this study directed the atypical pattern of phenotypic abnormalities. However, this may be unimportant for the purposes of drug and target discovery relating to cellular mechanisms of Tau toxicity, provided that the molecular mechanisms underlying neuronal dysfunction and death are conserved between the model and the disorder. Further work, including analysis of other zebrafish Tau lines constructed using different promoter elements, will be necessary to clarify this point. While detailed phenotypic analysis of the first cDNA transgenic Tau models is in progress, it is interesting to consider future directions in the generation of models and approaches for analysis. Use of models based on over-expression of cDNA encoding human Tau would limit discovery of novel molecular interventions to those targeting Tau modifications, Tau deposition or other abnormal effects exerted by Tau on neuronal function. However, there may be upstream targets related to transcription and splicing that would not be represented in cDNA models. For example, changes in the cellular 3R-/4R-Tau ratio, caused by alteration in the regulation of exon 10 splicing, underlie one form of FTDP17 and may contribute to sporadic Tauopathies, where associated haplotypes may alter 3R-/4R-Tau ratios [46]. It is not clear whether the human MAPT locus would be appropriately regulated in transgenic Tau zebrafish, allowing expression of physiological levels and patterns of alternatively spliced products. Transgenic zebrafish lines harboring the entire MAPT gene would therefore be of considerable interest. Furthermore, since zebrafish evolution may have resulted in the functions of 3R- and 4R-Tau isoforms becoming distributed between two different genes, an indication of whether alterations in the ratio of these forms can be pathogenic in zebrafish might be accomplished by morpholino knockdown experiments. Initial evidence shows that the zebrafish will most likely be a predictive platform for the discovery and assessment of Tau kinase inhibitors, although arguably the greatest potential strength of the model might be in discovering new therapeutic targets. In order to gain novel insights from screening approaches, end points that are independent of presumptions about underlying mechanisms, such as neurobehavioral analyses or cell death assays, might be deployed. Furthermore, experimental approaches that can be uniquely applied in zebrafish larvae owing to their optical transparency, for example physiological imaging modalities relating to cellular calcium levels, have the potential to yield important insights from hypothesisdriven experiments that cannot be carried out using other available models. In conclusion, the first reports show that construction of transgenic zebrafish expressing human Tau in CNS neurons is feasible and suggest that the resulting models will be relevant and useful. Consequently, there is cause for cautious optimism that novel zebrafish Tauopathy models may provide important contributions to ongoing efforts to address these common and devastating diseases. Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 Acknowledgements We gratefully acknowledge research grant support from CurePSP — The Society for Progressive Supranuclear Palsy (grants #441-05 and #468-08) and NINDS (NS058369). [25] [26] References [27] [1] M.D. Weingarten, A.H. Lockwood, S.Y. Hwo, M.W. Kirschner, A protein factor essential for microtubule assembly, Proc. Natl. Acad. Sci. U. S. A. 72 (1975) 1858–1862. [2] G.B. Witman, D.W. Cleveland, M.D. Weingarten, M.W. Kirschner, Tubulin requires tau for growth onto microtubule initiating sites, Proc. Natl. Acad. Sci. U. S. A. 73 (1976) 4070–4074. [3] L.I. Binder, A. Frankfurter, L.I. Rebhun, The distribution of tau in the mammalian central nervous system, J. Cell Biol. 101 (1985) 1371–1378. [4] K.A. Butner, M.W. Kirschner, Tau protein binds to microtubules through a flexible array of distributed weak sites, J. Cell Biol. 115 (1991) 717–730. [5] N. Gustke, B. Trinczek, J. Biernat, E.M. Mandelkow, E. Mandelkow, Domains of tau protein and interactions with microtubules, Biochemistry 33 (1994) 9511–9522. [6] M. Goedert, M.G. Spillantini, M.C. Potier, J. Ulrich, R.A. Crowther, Cloning and sequencing of the cDNA encoding an isoform of microtubule-associated protein tau containing four tandem repeats: differential expression of tau protein mRNAs in human brain, EMBO J. 8 (1989) 393–399. [7] M. Goedert, C.M. Wischik, R.A. Crowther, J.E. Walker, A. Klug, Cloning and sequencing of the cDNA encoding a core protein of the paired helical filament of Alzheimer disease: identification as the microtubule-associated protein tau, Proc. Natl. Acad. Sci. U. S. A. 85 (1988) 4051–4055. [8] M. Goedert, M.G. Spillantini, R. Jakes, D. Rutherford, R.A. Crowther, Multiple isoforms of human microtubule-associated protein tau: sequences and localization in neurofibrillary tangles of Alzheimer's disease, Neuron 3 (1989) 519–526. [9] M. Goedert, M.G. Spillantini, R.A. Crowther, Cloning of a big tau microtubuleassociated protein characteristic of the peripheral nervous system, Proc. Natl. Acad. Sci. U. S. A. 89 (1992) 1983–1987. [10] D.N. Drechsel, A.A. Hyman, M.H. Cobb, M.W. Kirschner, Modulation of the dynamic instability of tubulin assembly by the microtubule-associated protein tau, Mol. Biol. Cell 3 (1992) 1141–1154. [11] M. Goedert, R. Jakes, Expression of separate isoforms of human tau protein: correlation with the tau pattern in brain and effects on tubulin polymerization, EMBO J. 9 (1990) 4225–4230. [12] D. Panda, J.C. Samuel, M. Massie, S.C. Feinstein, L. Wilson, Differential regulation of microtubule dynamics by three- and four-repeat tau: implications for the onset of neurodegenerative disease, Proc. Natl. Acad. Sci. U. S. A. 100 (2003) 9548–9553. [13] B.L. Goode, S.C. Feinstein, Identification of a novel microtubule binding and assembly domain in the developmentally regulated inter-repeat region of tau, J. Cell Biol. 124 (1994) 769–782. [14] M. Hutton, C.L. Lendon, P. Rizzu, M. Baker, S. Froelich, H. Houlden, S. PickeringBrown, S. Chakraverty, A. Isaacs, A. Grover, J. Hackett, J. Adamson, S. Lincoln, D. Dickson, P. Davies, R.C. Petersen, M. Stevens, E. de Graaff, E. Wauters, J. van Baren, M. Hillebrand, M. Joosse, J.M. Kwon, P. Nowotny, P. Heutink, et al., Association of missense and 5′-splice-site mutations in tau with the inherited dementia FTDP17, Nature 393 (1998) 702–705. [15] R. Dixit, J.L. Ross, Y.E. Goldman, E.L. Holzbaur, Differential regulation of dynein and kinesin motor proteins by tau, Science 319 (2008) 1086–1089. [16] E.S. Matsuo, R.W. Shin, M.L. Billingsley, A. Van deVoorde, M. O'Connor, J.Q. Trojanowski, V.M. Lee, Biopsy-derived adult human brain tau is phosphorylated at many of the same sites as Alzheimer's disease paired helical filament tau, Neuron 13 (1994) 989–1002. [17] F. Liu, K. Iqbal, I. Grundke-Iqbal, G.W. Hart, C.X. Gong, O-GlcNAcylation regulates phosphorylation of tau: a mechanism involved in Alzheimer's disease, Proc. Natl. Acad. Sci. U. S. A. 101 (2004) 10804–10809. [18] G. Lee, R. Thangavel, V.M. Sharma, J.M. Litersky, K. Bhaskar, S.M. Fang, L.H. Do, A. Andreadis, G. Van Hoesen, H. Ksiezak-Reding, Phosphorylation of tau by fyn: implications for Alzheimer's disease, J. Neurosci. 24 (2004) 2304–2312. [19] V. Dorval, P.E. Fraser, Small ubiquitin-like modifier (SUMO) modification of natively unfolded proteins tau and alpha-synuclein, J. Biol. Chem. 281 (2006) 9919–9924. [20] T. Horiguchi, K. Uryu, B.I. Giasson, H. Ischiropoulos, R. LightFoot, C. Bellmann, C. Richter-Landsberg, V.M. Lee, J.Q. Trojanowski, Nitration of tau protein is linked to neurodegeneration in tauopathies, Am. J. Pathol. 163 (2003) 1021–1031. [21] M. Hasegawa, M. Morishima-Kawashima, K. Takio, M. Suzuki, K. Titani, Y. Ihara, Protein sequence and mass spectrometric analyses of tau in the Alzheimer's disease brain, J. Biol. Chem. 267 (1992) 17047–17054. [22] D.P. Hanger, B.H. Anderton, W. Noble, Tau phosphorylation: the therapeutic challenge for neurodegenerative disease, Trends Mol. Med. 15 (2009) 112–119. [23] J. Biernat, N. Gustke, G. Drewes, E.M. Mandelkow, E. Mandelkow, Phosphorylation of Ser262 strongly reduces binding of tau to microtubules: distinction between PHF-like immunoreactivity and microtubule binding, Neuron 11 (1993) 153–163. [24] N. Gustke, B. Steiner, E.M. Mandelkow, J. Biernat, H.E. Meyer, M. Goedert, E. Mandelkow, The Alzheimer-like phosphorylation of tau protein reduces [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] 361 microtubule binding and involves Ser-Pro and Thr-Pro motifs, FEBS Lett. 307 (1992) 199–205. P.J. Lu, G. Wulf, X.Z. Zhou, P. Davies, K.P. Lu, The prolyl isomerase Pin1 restores the function of Alzheimer-associated phosphorylated tau protein, Nature 399 (1999) 784–788. C.H. Reynolds, C.J. Garwood, S. Wray, C. Price, S. Kellie, T. Perera, M. Zvelebil, A. Yang, P.W. Sheppard, I.M. Varndell, D.P. Hanger, B.H. Anderton, Phosphorylation regulates tau interactions with Src homology 3 domains of phosphatidylinositol 3-kinase, phospholipase Cgamma1, Grb2, and Src family kinases, J. Biol. Chem. 283 (2008) 18177–18186. A. Kenessey, S.H. Yen, The extent of phosphorylation of fetal tau is comparable to that of PHF-tau from Alzheimer paired helical filaments, Brain Res. 629 (1993) 40–46. J.P. Brion, C. Smith, A.M. Couck, J.M. Gallo, B.H. Anderton, Developmental changes in tau phosphorylation: fetal tau is transiently phosphorylated in a manner similar to paired helical filament-tau characteristic of Alzheimer's disease, J. Neurochem. 61 (1993) 2071–2080. M. Hong, D.C. Chen, P.S. Klein, V.M. Lee, Lithium reduces tau phosphorylation by inhibition of glycogen synthase kinase-3, J. Biol. Chem. 272 (1997) 25326–25332. M. Hong, V.M. Lee, Insulin and insulin-like growth factor-1 regulate tau phosphorylation in cultured human neurons, J. Biol. Chem. 272 (1997) 19547–19553. W. Noble, E. Planel, C. Zehr, V. Olm, J. Meyerson, F. Suleman, K. Gaynor, L. Wang, J. LaFrancois, B. Feinstein, M. Burns, P. Krishnamurthy, Y. Wen, R. Bhat, J. Lewis, D. Dickson, K. Duff, Inhibition of glycogen synthase kinase-3 by lithium correlates with reduced tauopathy and degeneration in vivo, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 6990–6995. A. Caccamo, S. Oddo, L.X. Tran, F.M. LaFerla, Lithium reduces tau phosphorylation but not A beta or working memory deficits in a transgenic model with both plaques and tangles, Am. J. Pathol. 170 (2007) 1669–1675. D. Terwel, D. Muyllaert, I. Dewachter, P. Borghgraef, S. Croes, H. Devijver, F. Van Leuven, Amyloid activates GSK-3beta to aggravate neuronal tauopathy in bigenic mice, Am. J. Pathol. 172 (2008) 786–798. W. Noble, V. Olm, K. Takata, E. Casey, O. Mary, J. Meyerson, K. Gaynor, J. LaFrancois, L. Wang, T. Kondo, P. Davies, M. Burns, Veeranna, R. Nixon, D. Dickson, Y. Matsuoka, M. Ahlijanian, L.F. Lau, K. Duff, Cdk5 is a key factor in tau aggregation and tangle formation in vivo, Neuron 38 (2003) 555–565. J.C. Cruz, H.C. Tseng, J.A. Goldman, H. Shih, L.H. Tsai, Aberrant Cdk5 activation by p25 triggers pathological events leading to neurodegeneration and neurofibrillary tangles, Neuron 40 (2003) 471–483. K. Sobue, A. Agarwal-Mawal, W. Li, W. Sun, Y. Miura, H.K. Paudel, Interaction of neuronal Cdc2-like protein kinase with microtubule-associated protein tau, J. Biol. Chem. 275 (2000) 16673–16680. T. Li, C. Hawkes, H.Y. Qureshi, S. Kar, H.K. Paudel, Cyclin-dependent protein kinase 5 primes microtubule-associated protein tau site-specifically for glycogen synthase kinase 3beta, Biochemistry 45 (2006) 3134–3145. A. Sengupta, Q. Wu, I. Grundke-Iqbal, K. Iqbal, T.J. Singh, Potentiation of GSK-3catalyzed Alzheimer-like phosphorylation of human tau by cdk5, Mol. Cell. Biochem. 167 (1997) 99–105. M. Goedert, R. Jakes, Z. Qi, J.H. Wang, P. Cohen, Protein phosphatase 2A is the major enzyme in brain that dephosphorylates tau protein phosphorylated by proline-directed protein kinases or cyclic AMP-dependent protein kinase, J. Neurochem. 65 (1995) 2804–2807. E. Sontag, V. Nunbhakdi-Craig, G.S. Bloom, M.C. Mumby, A novel pool of protein phosphatase 2A is associated with microtubules and is regulated during the cell cycle, J. Cell Biol. 128 (1995) 1131–1144. M. Goedert, M.G. Spillantini, N.J. Cairns, R.A. Crowther, Tau proteins of Alzheimer paired helical filaments: abnormal phosphorylation of all six brain isoforms, Neuron 8 (1992) 159–168. N. Sergeant, J.P. David, D. Lefranc, P. Vermersch, A. Wattez, A. Delacourte, Different distribution of phosphorylated tau protein isoforms in Alzheimer's and Pick's diseases, FEBS Lett. 412 (1997) 578–582. T. Arai, K. Ikeda, H. Akiyama, Y. Shikamoto, K. Tsuchiya, S. Yagishita, T. Beach, J. Rogers, C. Schwab, P.L. McGeer, Distinct isoforms of tau aggregated in neurons and glial cells in brains of patients with Pick's disease, corticobasal degeneration and progressive supranuclear palsy, Acta Neuropathol. (Berl.) 101 (2001) 167–173. N. Sergeant, A. Wattez, A. Delacourte, Neurofibrillary degeneration in progressive supranuclear palsy and corticobasal degeneration: tau pathologies with exclusively “exon 10” isoforms, J. Neurochem. 72 (1999) 1243–1249. M. Baker, I. Litvan, H. Houlden, J. Adamson, D. Dickson, J. Perez-Tur, J. Hardy, T. Lynch, E. Bigio, M. Hutton, Association of an extended haplotype in the tau gene with progressive supranuclear palsy, Hum. Mol. Genet. 8 (1999) 711–715. T.M. Caffrey, C. Joachim, S. Paracchini, M.M. Esiri, R. Wade-Martins, Haplotypespecific expression of exon 10 at the human MAPT locus, Hum. Mol. Genet. 15 (2006) 3529–3537. M.G. Spillantini, J.R. Murrell, M. Goedert, M.R. Farlow, A. Klug, B. Ghetti, Mutation in the tau gene in familial multiple system tauopathy with presenile dementia, Proc. Natl. Acad. Sci. U. S. A. 95 (1998) 7737–7741. J. van Swieten, M.G. Spillantini, Hereditary frontotemporal dementia caused by Tau gene mutations, Brain Pathol. 17 (2007) 63–73. M. Hong, V. Zhukareva, V. Vogelsberg-Ragaglia, Z. Wszolek, L. Reed, B.I. Miller, D.H. Geschwind, T.D. Bird, D. McKeel, A. Goate, J.C. Morris, K.C. Wilhelmsen, G.D. Schellenberg, J.Q. Trojanowski, V.M. Lee, Mutation-specific functional impairments in distinct tau isoforms of hereditary FTDP-17, Science 282 (1998) 1914–1917. 362 Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 [50] V.M. Lee, T.K. Kenyon, J.Q. Trojanowski, Transgenic animal models of tauopathies, Biochim. Biophys. Acta 1739 (2005) 251–259. [51] F. Denk, R. Wade-Martins, Knock-out and transgenic mouse models of tauopathies, Neurobiol. Aging 30 (2009) 1–13. [52] Y. Yoshiyama, M. Higuchi, B. Zhang, S.M. Huang, N. Iwata, T.C. Saido, J. Maeda, T. Suhara, J.Q. Trojanowski, V.M. Lee, Synapse loss and microglial activation precede tangles in a P301S tauopathy mouse model, Neuron 53 (2007) 337–351. [53] K. Santacruz, J. Lewis, T. Spires, J. Paulson, L. Kotilinek, M. Ingelsson, A. Guimaraes, M. DeTure, M. Ramsden, E. McGowan, C. Forster, M. Yue, J. Orne, C. Janus, A. Mariash, M. Kuskowski, B. Hyman, M. Hutton, K.H. Ashe, Tau suppression in a neurodegenerative mouse model improves memory function, Science 309 (2005) 476–481. [54] T.L. Spires, J.D. Orne, K. SantaCruz, R. Pitstick, G.A. Carlson, K.H. Ashe, B.T. Hyman, Region-specific dissociation of neuronal loss and neurofibrillary pathology in a mouse model of tauopathy, Am. J. Pathol. 168 (2006) 1598–1607. [55] S. Oddo, V. Vasilevko, A. Caccamo, M. Kitazawa, D.H. Cribbs, F.M. LaFerla, Reduction of soluble Abeta and tau, but not soluble Abeta alone, ameliorates cognitive decline in transgenic mice with plaques and tangles, J. Biol. Chem. 281 (2006) 39413–39423. [56] J. Gotz, F. Chen, J. van Dorpe, R.M. Nitsch, Formation of neurofibrillary tangles in P301l tau transgenic mice induced by Abeta 42 fibrils, Science 293 (2001) 1491–1495. [57] J. Lewis, D.W. Dickson, W.L. Lin, L. Chisholm, A. Corral, G. Jones, S.H. Yen, N. Sahara, L. Skipper, D. Yager, C. Eckman, J. Hardy, M. Hutton, E. McGowan, Enhanced neurofibrillary degeneration in transgenic mice expressing mutant tau and APP, Science 293 (2001) 1487–1491. [58] K.R. Brunden, J.Q. Trojanowski, V.M. Lee, Advances in tau-focused drug discovery for Alzheimer's disease and related tauopathies, Nat. Rev. Drug Discov. 8 (2009) 783–793. [59] S. Le Corre, H.W. Klafki, N. Plesnila, G. Hubinger, A. Obermeier, H. Sahagun, B. Monse, P. Seneci, J. Lewis, J. Eriksen, C. Zehr, M. Yue, E. McGowan, D.W. Dickson, M. Hutton, H.M. Roder, An inhibitor of tau hyperphosphorylation prevents severe motor impairments in tau transgenic mice, Proc. Natl. Acad. Sci. U. S. A. 103 (2006) 9673–9678. [60] B. Zhang, A. Maiti, S. Shively, F. Lakhani, G. McDonald-Jones, J. Bruce, E.B. Lee, S.X. Xie, S. Joyce, C. Li, P.M. Toleikis, V.M. Lee, J.Q. Trojanowski, Microtubule-binding drugs offset tau sequestration by stabilizing microtubules and reversing fast axonal transport deficits in a tauopathy model, Proc. Natl. Acad. Sci. U. S. A. 102 (2005) 227–231. [61] C.A. Dickey, A. Kamal, K. Lundgren, N. Klosak, R.M. Bailey, J. Dunmore, P. Ash, S. Shoraka, J. Zlatkovic, C.B. Eckman, C. Patterson, D.W. Dickson, N.S. Nahman Jr., M. Hutton, F. Burrows, L. Petrucelli, The high-affinity HSP90 – CHIP complex recognizes and selectively degrades phosphorylated tau client proteins, J. Clin. Invest. 117 (2007) 648–658. [62] K. Kawakami, Transgenesis and gene trap methods in zebrafish by using the Tol2 transposable element, Methods Cell Biol. 77 (2004) 201–222. [63] D. Soroldoni, B.M. Hogan, A.C. Oates, Simple and efficient transgenesis with meganuclease constructs in zebrafish, Methods Mol. Biol. 546 (2009) 117–130. [64] A. Nasevicius, S.C. Ekker, Effective targeted gene ‘knockdown’ in zebrafish, Nat. Genet. 26 (2000) 216–220. [65] K.G. Peters, P.S. Rao, B.S. Bell, L.A. Kindman, Green fluorescent fusion proteins: powerful tools for monitoring protein expression in live zebrafish embryos, Dev. Biol. 171 (1995) 252–257. [66] D.L. McLean, J.R. Fetcho, Using imaging and genetics in zebrafish to study developing spinal circuits in vivo, Dev. Neurobiol. 68 (2008) 817–834. [67] L. Wen, W. Wei, W. Gu, P. Huang, X. Ren, Z. Zhang, Z. Zhu, S. Lin, B. Zhang, Visualization of monoaminergic neurons and neurotoxicity of MPTP in live transgenic zebrafish, Dev. Biol. 314 (2008) 84–92. [68] F. Peri, C. Nusslein-Volhard, Live imaging of neuronal degradation by microglia reveals a role for v0-ATPase a1 in phagosomal fusion in vivo, Cell 133 (2008) 916–927. [69] W. Driever, L. Solnica-Krezel, A.F. Schier, S.C. Neuhauss, J. Malicki, D.L. Stemple, D.Y. Stainier, F. Zwartkruis, S. Abdelilah, Z. Rangini, J. Belak, C. Boggs, A genetic screen for mutations affecting embryogenesis in zebrafish, Development 123 (1996) 37–46. [70] L. Solnica-Krezel, A.F. Schier, W. Driever, Efficient recovery of ENU-induced mutations from the zebrafish germline, Genetics 136 (1994) 1401–1420. [71] G. Golling, A. Amsterdam, Z. Sun, M. Antonelli, E. Maldonado, W. Chen, S. Burgess, M. Haldi, K. Artzt, S. Farrington, S.Y. Lin, R.M. Nissen, N. Hopkins, Insertional mutagenesis in zebrafish rapidly identifies genes essential for early vertebrate development, Nat. Genet. 31 (2002) 135–140. [72] A. Amsterdam, S. Burgess, G. Golling, W. Chen, Z. Sun, K. Townsend, S. Farrington, M. Haldi, N. Hopkins, A large-scale insertional mutagenesis screen in zebrafish, Genes Dev. 13 (1999) 2713–2724. [73] L.I. Zon, R.T. Peterson, In vivo drug discovery in the zebrafish, Nat. Rev. Drug Discov. 4 (2005) 35–44. [74] H.M. Stern, R.D. Murphey, J.L. Shepard, J.F. Amatruda, C.T. Straub, K.L. Pfaff, G. Weber, J.A. Tallarico, R.W. King, L.I. Zon, Small molecules that delay S phase suppress a zebrafish bmyb mutant, Nat. Chem. Biol. 1 (2005) 366–370. [75] G. Molina, A. Vogt, A. Bakan, W. Dai, P. Queiroz de Oliveira, W. Znosko, T.E. Smithgall, I. Bahar, J.S. Lazo, B.W. Day, M. Tsang, Zebrafish chemical screening reveals an inhibitor of Dusp6 that expands cardiac cell lineages, Nat. Chem. Biol. 5 (2009) 680–687. [76] J. Rihel, D.A. Prober, A. Arvanites, K. Lam, S. Zimmerman, S. Jang, S.J. Haggarty, D. Kokel, L.L. Rubin, R.T. Peterson, A.F. Schier, Zebrafish behavioral profiling links drugs to biological targets and rest/wake regulation, Science 327 (2010) 348–351. [77] M.F. Wullimann, B. Rupp, H. Reichert, Neuroanatomy of the Zebrafish Brain, Birkhauser-Verlag, Berlin, 1996. [78] T. Mueller, M.F. Wullimann, An evolutionary interpretation of teleostean forebrain anatomy, Brain Behav. Evol. 74 (2009) 30–42. [79] E. Rink, M.F. Wullimann, Connections of the ventral telencephalon (subpallium) in the zebrafish (Danio rerio), Brain Res. 1011 (2004) 206–220. [80] T. Mueller, P. Vernier, M.F. Wullimann, The adult central nervous cholinergic system of a neurogenetic model animal, the zebrafish Danio rerio, Brain Res. 1011 (2004) 156–169. [81] V. Sallinen, V. Torkko, M. Sundvik, I. Reenilä, D. Khrustalyov, J. Kaslin, P. Panula, MPTP and MPP+ target specific aminergic cell populations in larval zebrafish, J. Neurochem. 108 (3) (2008) 719–731. [82] R.L. Bernardos, P.A. Raymond, GFAP transgenic zebrafish, Gene Expr. Patterns 6 (2006) 1007–1013. [83] K. Tomizawa, Y. Inoue, H. Nakayasu, A monoclonal antibody stains radial glia in the adult zebrafish (Danio rerio) CNS, J. Neurocytol. 29 (2000) 119–128. [84] B.B. Kirby, N. Takada, A.J. Latimer, J. Shin, T.J. Carney, R.N. Kelsh, B. Appel, In vivo time-lapse imaging shows dynamic oligodendrocyte progenitor behavior during zebrafish development, Nat. Neurosci. 9 (2006) 1506–1511. [85] M. Yoshida, W.B. Macklin, Oligodendrocyte development and myelination in GFP-transgenic zebrafish, J. Neurosci. Res. 81 (2005) 1–8. [86] R.L. Avila, B.R. Tevlin, J.P. Lees, H. Inouye, D.A. Kirschner, Myelin structure and composition in zebrafish, Neurochem. Res. 32 (2007) 197–209. [87] J. Schweitzer, T. Becker, M. Schachner, K.A. Nave, H. Werner, Evolution of myelin proteolipid proteins: gene duplication in teleosts and expression pattern divergence, Mol. Cell. Neurosci. 31 (2006) 161–177. [88] J.Y. Jeong, H.B. Kwon, J.C. Ahn, D. Kang, S.H. Kwon, J.A. Park, K.W. Kim, Functional and developmental analysis of the blood – brain barrier in zebrafish, Brain Res. Bull. 75 (2008) 619–628. [89] S. Bretaud, C. Allen, P.W. Ingham, O. Bandmann, p53-dependent neuronal cell death in a DJ-1-deficient zebrafish model of Parkinson's disease, J. Neurochem. 100 (2007) 1626–1635. [90] Q. Bai, S.J. Mullett, J.A. Garver, D.A. Hinkle, E.A. Burton, Zebrafish DJ-1 is evolutionarily conserved and expressed in dopaminergic neurons, Brain Res. 1113 (2006) 33–44. [91] O. Anichtchik, H. Diekmann, A. Fleming, A. Roach, P. Goldsmith, D.C. Rubinsztein, Loss of PINK1 function affects development and results in neurodegeneration in zebrafish, J. Neurosci. 28 (2008) 8199–8207. [92] Z. Sun, A.D. Gitler, Discovery and characterization of three novel synuclein genes in zebrafish, Dev. Dyn. 237 (2008) 2490–2495. [93] L. Flinn, H. Mortiboys, K. Volkmann, R.W. Koster, P.W. Ingham, O. Bandmann, Complex I deficiency and dopaminergic neuronal cell loss in parkin-deficient zebrafish (Danio rerio), Brain 132 (2009) 1613–1623. [94] W.A. Campbell, H. Yang, H. Zetterberg, S. Baulac, J.A. Sears, T. Liu, S.T. Wong, T.P. Zhong, W. Xia, Zebrafish lacking Alzheimer presenilin enhancer 2 (Pen-2) demonstrate excessive p53-dependent apoptosis and neuronal loss, J. Neurochem. 96 (2006) 1423–1440. [95] C. Groth, S. Nornes, R. McCarty, R. Tamme, M. Lardelli, Identification of a second presenilin gene in zebrafish with similarity to the human Alzheimer's disease gene presenilin2, Dev. Genes Evol. 212 (2002) 486–490. [96] U. Leimer, K. Lun, H. Romig, J. Walter, J. Grunberg, M. Brand, C. Haass, Zebrafish (Danio rerio) presenilin promotes aberrant amyloid beta-peptide production and requires a critical aspartate residue for its function in amyloidogenesis, Biochemistry 38 (1999) 13602–13609. [97] M. Chen, R.N. Martins, M. Lardelli, Complex splicing and neural expression of duplicated tau genes in zebrafish embryos, J. Alzheimers Dis. 18 (2) (2009) 305–317. [98] A. Amores, A. Force, Y.L. Yan, L. Joly, C. Amemiya, A. Fritz, R.K. Ho, J. Langeland, V. Prince, Y.L. Wang, M. Westerfield, M. Ekker, J.H. Postlethwait, Zebrafish hox clusters and vertebrate genome evolution, Science 282 (1998) 1711–1714. [99] Y.C. Chen, M. Priyadarshini, P. Panula, Complementary developmental expression of the two tyrosine hydroxylase transcripts in zebrafish, Histochem. Cell Biol. 132 (4) (2009) 375–381. [100] G.W. Stuart, J.V. McMurray, M. Westerfield, Replication, integration and stable germ-line transmission of foreign sequences injected into early zebrafish embryos, Development 103 (1988) 403–412. [101] A. Jacquier, B. Dujon, An intron-encoded protein is active in a gene conversion process that spreads an intron into a mitochondrial gene, Cell 41 (1985) 383–394. [102] V. Thermes, C. Grabher, F. Ristoratore, F. Bourrat, A. Choulika, J. Wittbrodt, J.S. Joly, I-SceI meganuclease mediates highly efficient transgenesis in fish, Mech. Dev. 118 (2002) 91–98. [103] Q. Bai, X. Wei, E.A. Burton, Expression of a 12-kb promoter element derived from the zebrafish enolase-2 gene in the zebrafish visual system, Neurosci. Lett. 449 (2009) 252–257. [104] A. Koga, M. Suzuki, H. Inagaki, Y. Bessho, H. Hori, Transposable element in fish, Nature 383 (1996) 30. [105] K. Kawakami, A. Shima, N. Kawakami, Identification of a functional transposase of the Tol2 element, an Ac-like element from the Japanese medaka fish, and its transposition in the zebrafish germ lineage, Proc. Natl. Acad. Sci. U. S. A. 97 (2000) 11403–11408. [106] S. Higashijima, H. Okamoto, N. Ueno, Y. Hotta, G. Eguchi, High-frequency generation of transgenic zebrafish which reliably express GFP in whole muscles or the whole body by using promoters of zebrafish origin, Dev. Biol. 192 (1997) 289–299. [107] Q. Long, A. Meng, H. Wang, J.R. Jessen, M.J. Farrell, S. Lin, GATA-1 expression pattern can be recapitulated in living transgenic zebrafish using GFP reporter gene, Development 124 (1997) 4105–4111. Q. Bai, E.A. Burton / Biochimica et Biophysica Acta 1812 (2011) 353–363 [108] A. Meng, H. Tang, B.A. Ong, M.J. Farrell, S. Lin, Promoter analysis in living zebrafish embryos identifies a cis-acting motif required for neuronal expression of GATA-2, Proc. Natl. Acad. Sci. U. S. A. 94 (1997) 6267–6272. [109] C.H. Kim, E. Ueshima, O. Muraoka, H. Tanaka, S.Y. Yeo, T.L. Huh, N. Miki, Zebrafish elav/HuC homologue as a very early neuronal marker, Neurosci. Lett. 216 (1996) 109–112. [110] H.C. Park, C.H. Kim, Y.K. Bae, S.Y. Yeo, S.H. Kim, S.K. Hong, J. Shin, K.W. Yoo, M. Hibi, T. Hirano, N. Miki, A.B. Chitnis, T.L. Huh, Analysis of upstream elements in the HuC promoter leads to the establishment of transgenic zebrafish with fluorescent neurons, Dev. Biol. 227 (2000) 279–293. [111] Q. Bai, J.A. Garver, N.A. Hukriede, E.A. Burton, Generation of a transgenic zebrafish model of Tauopathy using a novel promoter element derived from the zebrafish eno2 gene, Nucleic Acids Res. 35 (2007) 6501–6516. [112] K. Asakawa, M.L. Suster, K. Mizusawa, S. Nagayoshi, T. Kotani, A. Urasaki, Y. Kishimoto, M. Hibi, K. Kawakami, Genetic dissection of neural circuits by Tol2 transposon-mediated Gal4 gene and enhancer trapping in zebrafish, Proc. Natl. Acad. Sci. U. S. A. 105 (2008) 1255–1260. [113] M. Distel, M.F. Wullimann, R.W. Koster, Optimized Gal4 genetics for permanent gene expression mapping in zebrafish, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 13365–13370. [114] D. Paquet, R. Bhat, A. Sydow, E.M. Mandelkow, S. Berg, S. Hellberg, J. Falting, M. Distel, R.W. Koster, B. Schmid, C. Haass, A zebrafish model of tauopathy allows in vivo imaging of neuronal cell death and drug evaluation, J. Clin. Invest. 119 (2009) 1382–1395. [115] E.K. Scott, L. Mason, A.B. Arrenberg, L. Ziv, N.J. Gosse, T. Xiao, N.C. Chi, K. Asakawa, K. Kawakami, H. Baier, Targeting neural circuitry in zebrafish using GAL4 enhancer trapping, Nat. Methods 4 (2007) 323–326. [116] D. Dickson, Sporadic tauopathies: Pick's disease, corticobasal degeneration, progressive supranuclear palsy and argyrophilic grain disease, in: M. Esiri, V.M. Lee, J.Q. Trojanowski (Eds.), The Neuropathology of Dementia, Cambridge University Press, Cambridge, 2004, pp. 227–256. 363 [117] Y. Gao, P. Li, L. Li, Transgenic zebrafish that express tyrosine hydroxylase promoter in inner retinal cells, Dev. Dyn. 233 (2005) 921–929. [118] S. Meng, S. Ryu, B. Zhao, D.Q. Zhang, W. Driever, D.G. McMahon, Targeting retinal dopaminergic neurons in tyrosine hydroxylase-driven green fluorescent protein transgenic zebrafish, Mol. Vis. 14 (2008) 2475–2483. [119] Q. Bai, E.A. Burton, Cis-acting elements responsible for dopaminergic neuronspecific expression of zebrafish slc6a3 (dopamine transporter) in vivo are located remote from the transcriptional start site, Neuroscience 164 (2009) 1138–1151. [120] O.V. Anichtchik, J. Kaslin, N. Peitsaro, M. Scheinin, P. Panula, Neurochemical and behavioural changes in zebrafish Danio rerio after systemic administration of 6hydroxydopamine and 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine, J. Neurochem. 88 (2004) 443–453. [121] H.G. Tomasiewicz, D.B. Flaherty, J.P. Soria, J.G. Wood, Transgenic zebrafish model of neurodegeneration, J. Neurosci. Res. 70 (2002) 734–745. [122] G.A. Jicha, R. Bowser, I.G. Kazam, P. Davies, Alz-50 and MC-1, a new monoclonal antibody raised to paired helical filaments, recognize conformational epitopes on recombinant tau, J. Neurosci. Res. 48 (1997) 128–132. [123] M. Kidd, Paired helical filaments in electron microscopy of Alzheimer's disease, Nature 197 (1963) 192–193. [124] I. Tellez-Nagel, H.M. Wisniewski, Ultrastructure of neurofibrillary tangles in Steele–Richardson–Olszewski syndrome, Arch. Neurol. 29 (1973) 324–327. [125] H. Ksiezak-Reding, K. Morgan, L.A. Mattiace, P. Davies, W.K. Liu, S.H. Yen, K. Weidenheim, D.W. Dickson, Ultrastructure and biochemical composition of paired helical filaments in corticobasal degeneration, Am. J. Pathol. 145 (1994) 1496–1508. [126] D. Munoz-Garcia, S.K. Ludwin, Classic and generalized variants of Pick's disease: a clinicopathological, ultrastructural, and immunocytochemical comparative study, Ann. Neurol. 16 (1984) 467–480. [127] F. Gallyas, Silver staining of Alzheimer's neurofibrillary changes by means of physical development, Acta Morphol. Acad. Sci. Hung. 19 (1971) 1–8. Biochimica et Biophysica Acta 1812 (2011) 364–380 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Investigating regeneration and functional integration of CNS neurons: Lessons from zebrafish genetics and other fish species☆ Valerie C. Fleisch a,b,⁎, Brittany Fraser b, W. Ted Allison a,b,c a b c Centre for Prions & Protein Folding Disease, University of Alberta, Edmonton, Alberta, T6G 2M8, Canada Department of Biological Sciences, University of Alberta, Edmonton, Alberta, T6G 2J9, Canada Department of Medical Genetics, University of Alberta, Edmonton, Alberta, T6G 2H7, Canada a r t i c l e i n f o Article history: Received 6 January 2010 Received in revised form 5 October 2010 Accepted 21 October 2010 Available online 28 October 2010 Keywords: Zebrafish Regeneration Retina Optic nerve Spinal cord Brain Central nervous system Stem cell Glia a b s t r a c t Zebrafish possess a robust, innate CNS regenerative ability. Combined with their genetic tractability and vertebrate CNS architecture, this ability makes zebrafish an attractive model to gain requisite knowledge for clinical CNS regeneration. In treatment of neurological disorders, one can envisage replacing lost neurons through stem cell therapy or through activation of latent stem cells in the CNS. Here we review the evidence that radial glia are a major source of CNS stem cells in zebrafish and thus activation of radial glia is an attractive therapeutic target. We discuss the regenerative potential and the molecular mechanisms thereof, in the zebrafish spinal cord, retina, optic nerve and higher brain centres. We evaluate various cell ablation paradigms developed to induce regeneration, with particular emphasis on the need for (high throughput) indicators that neuronal regeneration has restored sensory or motor function. We also examine the potential confound that regeneration imposes as the community develops zebrafish models of neurodegeneration. We conclude that zebrafish combine several characters that make them a potent resource for testing hypotheses and discovering therapeutic targets in functional CNS regeneration. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Introduction It had long been assumed that the mammalian central nervous system (CNS) is incapable of regenerating neurons. However, today it is widely accepted that sources of adult neurogenesis do exist, and that these areas have the potential to underpin regeneration. Adult neurogenesis is localized to the dentate gyrus of the hippocampus as well as the lateral ventricles of the mammalian forebrain [1–3]. It has been shown that injury to the CNS (caused by ischemia, seizures, radiation or neurodegenerative disorders) triggers proliferation of progenitor cells in the subgranular zone and the subventricular zone in mammalian animal models and humans. However, in most cases these newborn cells do not show long-term survival, fail to migrate or integrate properly into the neuronal network (for a review see [4]). This lack of regenerative ability seems to have arisen in the course of evolution; most invertebrates (commonly used models are Hydra and planarians) as well as a number of anamniotic vertebrates, such as amphibians and fish, are capable of regenerating whole body parts and organs (for a review see [5]). Although invertebrates probably show ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Corresponding author. Centre for Prions & Protein Folding Disease, University of Alberta, Edmonton, Alberta, T6G 2M8, Canada. Tel.: +1 780 492 1087. E-mail address: valerie.fl[email protected] (V.C. Fleisch). 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.10.012 the highest degree of regenerative power, lack of genetic tractability as well as the absence of higher level brain centres are major drawbacks for their use as animal models in CNS regeneration research. Among vertebrates, a similarly high regenerative potential is found in amphibians, such as the newt or Xenopus, and in fish. Newts are extremely difficult to maintain and breed under laboratory conditions, and genetic tools have only been developed recently. Xenopus and zebrafish are both popular animal models for CNS regeneration research, but this review will focus on the teleost model zebrafish. Zebrafish are very potent in regenerating a variety of tissues. Traditionally, the zebrafish has been used as a model for studying the regeneration of the fins [6]. Since the retina is the most easily accessible part of the CNS, and morphology and function of the retina are basically conserved among vertebrate species, zebrafish have also been extensively used for studying the regeneration of the retina and optic nerve (reviewed below). Moreover, zebrafish have also increasingly been used to study regeneration of other nervous tissues, such as brain nuclei and the spinal cord. In contrast to newts, zebrafish are very easy and cost-effective to breed and produce high numbers of offspring [7]. Moreover, the abundance of genetic tools, including transgenic lines, knock-down (morpholino antisense technology [8]), mutants and novel knock-out (zinc finger nuclease [9,10]) techniques and microarrays, facilitate research aimed at identifying molecular pathways underlying regenerative processes. Access to zebrafish embryos is easy, as larvae develop outside the mother, and V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 transparency of larvae enables in vivo tracking of regenerating cells. Zebrafish also offer very powerful behavioral and electrophysiological tools to assess functional integration of regenerated neurons/axons [11,12]. Studying the exact mechanisms of how anamniotes such as zebrafish utilize stem cells to replace neurons of the CNS is valuable for helping us to understand why humans/mammals mostly lack this regenerative ability and more importantly, will provide us with ideas of how human CNS tissue could be stimulated to regenerate after injury. This type of knowledge will be crucial for advancing stem cell therapeutics. As direct transplantation of stem cells has been proven inefficient for restoring missing/damaged tissues, future approaches might involve switching existing dormant progenitor cells or even terminally differentiated cells back to a “stem cell mode”. In order to address this scientific puzzle, a number of fundamental questions immediately come to mind: What is the identity of the stem cells giving rise to newborn neurons — are these cells the same ones across species and across different tissues? Is the limitation on CNS regeneration intrinsic to the neuron or do extrinsic signals control regeneration (role of glia cell in inhibition versus promotion of regeneration)? What are the molecular pathways underlying dedifferentiation of glial stem cells and specification of new neurons? Are regenerated neurons functionally integrated into the CNS and how is that achieved? We are beginning to address these questions in zebrafish models of CNS regeneration, and the present review aims to provide a summary of the (partial) answers that have arisen from these studies. 2. Regeneration of various CNS tissues Studies of neural regeneration in zebrafish have been carried out in various CNS tissues. We review each in turn, dividing the topics into retina, optic nerve, spinal cord, and brain nuclei. 2.1. Retina Ocular disorders involving degeneration of cells in the neural retina are exceedingly common. Death of rod and/or cone photoreceptors is found in retinal dystrophies, such as AMD (age-related macular degeneration), Retinitis Pigmentosa, Usher Syndrome or Leber Congenital Amaurosis. Death of retinal ganglion cells is the largest contributor to vision loss in glaucoma, a condition affecting the optic nerve. Cells lost from the retina cannot be replaced, as the mammalian retina does not have intrinsic repair processes, rendering the patient severely visually impaired or even completely blind. In contrast to mammals, anamniotic vertebrates like zebrafish have amazing regenerative capabilities, including the capacity to regenerate any cell type in the retina. Cells are continuously added to the zebrafish retina as it expands throughout the life of the fish. This unique feature of teleost fish requires that sources of neurogenesis are maintained even in the adult. In the intact retina, progenitor cells have been classified as being localized to two distinct cellular compartments [13,14]. The CMZ (ciliary marginal zone) gives rise to all retinal cell types, whereas rod photoreceptors arise from a separate lineage of CMZ-derived Müller glia [13,15,16]. Neuronal progenitors of the CMZ express a number of markers such as n-cadherin, components of the Notch–Delta signalling pathway, rx1, vsx2/chx10 and pax6a [16]. Proliferating Müller glia cells in the INL (inner nuclear layer) produce rod progenitors which subsequently migrate to the ONL (outer nuclear layer), where they differentiate into mature rod photoreceptors [15,17,18]. Cells within both of these compartments show characteristic features of stem cells, including asymmetric division and self-renewal. An additional population of stem cells is activated upon retinal injury and subsequently produces progenitors for regenerating lost retinal cells (for a review see [16,19]). It has been shown that in fish, damage to the retina induces the formation of 365 columns of proliferating cells, termed “neurogenic clusters”, which express a similar set of stem cell markers as progenitor cells residing in the CMZ such as pax6, vsx1, notch-3 and n-cadherin [20–23]. Vihtelic et al. showed that PCNA (proliferating cell nuclear antigen, a marker for proliferation) positive cells can be detected as early as one day after light-induced photoreceptor cell death in the INL (inner nuclear layer) in albino zebrafish. These progenitor cells migrate to the ONL (outer nuclear layer) within 96 h, where they differentiate into rod and cone photoreceptors [24,25]. The identity of the cells comprising these clusters was resolved only recently when several researchers found evidence that Müller glia are the major source of stem cells in the adult zebrafish retina. 2.1.1. Retinal astrocytes (Müller glia cells) function as stem cells in the regenerating adult zebrafish retina Studies in recent years have dramatically changed our view of astrocytes as “mere support cells” or “neuronal glue”. Astrocytes are now viewed as highly important cells with a number of complex functions. Astrocytes are found in all tissues of the CNS (e.g. cerebellar Bergmann glia, retinal Müller glia) and have a number of established roles in metabolic support for neurons, K+ and neurotransmitter uptake, synaptogenesis, angiogenesis and maintenance of the brain– blood barrier (for a detailed recent review see [26]). Groundbreaking studies of neuronal precursors in the CNS have elucidated the existence of yet another type of glia cell [27,28]. These cells, exhibiting neuronal stem cell features, are located in adult neurogenic regions (the adult SVZ, subventricular zone and SGZ, subgranular zone). They have been shown to give rise to newborn olfactory neurons [29–31] and granule neurons [32–34]. The brain is not the only CNS tissue containing a large number of astrocytes. Müller glia of the retina perform similar functions as other glial cells in the brain, including structural and metabolic support of neurons, ion buffering, and neurotransmitter homeostasis. Studies in the chick retina [35,36] as well as experiments using transgenic zebrafish lines in combination with lineage tracing have revealed that, in addition to these basic support functions, Müller glia cells act as retinal stem cells after injury. Following photoreceptor cell death, they undergo morphological as well as gene expression changes. Thummel et al. have shown that Müller glia cells dedifferentiate upon injury as they lose expression of cell-type-specific markers such as GFAP (glial fibrillary acidic protein) and GS (glutamine synthetase) [37]. Signs of dedifferentiation include re-expression of BLBP (brain lipid binding protein), up-regulation of components of the Notch–Delta signalling pathway and redistribution of N-cadherin to the basolateral plasma membrane [16]. In succession, several groups used lineage tracing via transgenic expression of GFP in Müller cells in combination with immunocytological stainings and BrdU labelling to provide evidence that Müller cells not only respond to injury, but also contribute to the stem cells of the retina and give rise to neuronal progenitors (Fig. 1). Fausett and Goldman used the α1T transgenic line, driving transgene expression in a subpopulation of proliferative Müller glia cells after retinal injury [38]. BrdU labelling experiments revealed a dramatic increase of BrdU+ cells in the INL at 2 dpi (days post-injury), preceding a rise in BrdU+ cells in the ONL at 3 dpi. Most of these proliferating cells were GFP+, had a Müller cell-like morphology and expressed the Müller cell markers GFAP and GS. BrdU labelling at a later time point, 2 dpi, revealed staining in the ONL and ganglion cell layer (GCL), suggesting migration of INL precursors. The progenitors that had presumably arisen from reactive Müller glia cells started to express neuron-specific markers after 7 dpi, and were localized to the INL, IPL and GCL. BrdU addition at 4 dpi and evaluation of cells at 60 and 180 dpi revealed cells expressing markers of bipolar and amacrine cells, Müller glia and cone photoreceptors [39]. Bernardos et al. used a gfap:GFP transgenic zebrafish line harbouring the Müller cell-specific gfap promoter driving GFP expression specifically in Müller glia cells [40]. Following light 366 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 Fig. 1. Radial glia throughout the CNS are a source of stem cells. Here we depict retinal Müller glia cells as an example of radial glial cells of the CNS. The retina of a 4 dpf zebrafish larva with its three nuclear layers (GCL, ganglion cell layer; INL, inner nuclear layer; ONL, outer nuclear layer) is shown as a reference for the location of retinal progenitors (A). Differentiated radial glia cells spanning the larval (and adult) zebrafish retina (B) become activated upon injury, dedifferentiate and start to proliferate (C). Neuronal progenitors of all retinal cell lineages arise from these activated cells (D), and migrate to specific retinal layers where they differentiate into neurons (E). damage, the number of GFP+ Müller cells increased 3 fold up to 5 dpi, indicating a proliferative response. Coinciding with a peak in GFP+ cells, the expression of the progenitor cell markers pax6, pcna and crx (a marker for photoreceptors progenitors) peaked between 1 and 5 dpi. The spatio-temporal distribution of pax6+ and crx+ cells was consistent with a transition from pax6 to crx expression in progenitors as they migrated from the INL to the ONL. After 1 month, cones and rods regenerated completely and Müller cell density fell back to unlesioned levels. Morris et al. specifically induced death of cone photoreceptors by using the pde6cw59 mutant zebrafish line, harbouring a mutation in the cone-specific phosphodiesterase gene. BrdU exposure 4 h before sacrifice and subsequent CAZ (carbonic anhydrase, a marker for Müller glia cells) staining revealed a significant increase of BrdU+ cells in the INL, partly co-localizing with CAZ. These results suggest that at least a subset of Müller glia cells had re-entered the cell cycle [41]. Studies using morpholino-mediated knock-down of PCNA [42] provided evidence that progenitor cell formation in zebrafish retina after injury is dependent on proliferation of Müller glia cells. PCNA knock-down prior to light damage inhibited proliferation of Müller glia cells and resulted in decreased levels of GS expression as well as Müller cell death. Importantly, rod photoreceptors and short and long single cones could not be regenerated in morpholino-injected zebrafish larvae [42]. Assuming the effects of PCNA knock-down are primarily within the Müller glia, these data strongly supports the hypothesis that these cells are the source of retina stem cells. Several researchers showed that damage to photoreceptors triggers a regenerative response through activation of Müller glia cells. Initially, researchers had speculated that apoptotic photoreceptors would provide a molecular signal to Müller glia cells, thereby activating dedifferentiation and proliferation. However, Fimbel et al. showed that damage to the GCL and INL, without substantial damage to photoreceptors, likewise causes Müller cells to re-enter the cell cycle and generate retinal progenitors [43]. Low doses of intravitreal ouabain injections caused GCL and INL death. At 1 day post-injury (dpi), PCNA expression was seen in Müller glia cells, and by 5 dpi, its expression shifted to neuronal progenitors expressing the olig2:egfp transgene. After 60 dpi, 75% of ganglion and amacrine cells had regenerated [43]. Various studies support Müller glia cells as one source, perhaps even a major source, of progenitor cells in the zebrafish retina. Müller glia cells are activated following retinal damage and subsequently dedifferentiate, proliferate and give rise to different retinal cell types (Fig. 1). As Müller glia cells are also found in the mammalian/human retina, understanding which molecular cues are required to turn a quiescent Müller glia cell into a retinal stem cell will be crucial for the development of potential novel therapeutics which could be applied to trigger regeneration in the human retina. Although under normal physiological conditions, mammalian Müller glia cells do not re-enter the cell cycle in response to injury [44–46], it has been speculated that the human retina may have some regenerative potential. Several studies have reported the presence of BrdU+ photoreceptors after retina damage or after the addition of growth factors [47–51]. Moreover, the human retina may contain a compartment similar to the zebrafish CMZ. In retinal explants, cells of the human retinal margin as well as the INL re-enter the cell cycle when EGF (epidermal V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 growth factor) is added to the culture medium [52]. Thus the human retina may harbour quiescent stem cells, which can be triggered to reenter the cell cycle by the addition of growth factors. Unfortunately very little is known about the molecular pathways leading to the cell-biological changes inducing the activation of Müller glia cells. In recent years, a number of microarray studies focussing on temporal expression changes during retina regeneration in zebrafish have been conducted and have identified several candidate genes and pathways presumably implicated in the regeneration process. Data collected in these studies will be presented in the next section of this review, and an overview of the molecular components that have been identified can be found in Table 1. 2.1.2. Molecular pathways involved in retinal regeneration In order to identify candidate genes involved in the regenerative response in the zebrafish retina, a number of microarray-based studies have been performed. While one study used surgical removal of a small piece of the retina [53], mostly light damage paradigms have been applied to cause retinal injury [54–57]. Microarray analysis conducted on mRNA isolated from whole retina showed significant changes in expression of genes implicated in the cell cycle, cell proliferation, cell death, DNA replication, general metabolism and axonal regeneration as expected [37,53,54,56]. Genes encoding the opsin proteins were decreased in expression, reflecting death of photoreceptors, whereas cell cycle genes increased their expression, corresponding to progenitor proliferation [54]. As regenerationspecific responses are expected to occur in specific cell types (Müller glia cells and emanating progenitor cells) expression analysis of RNA extracted exclusively from those cells was expected to reveal novel regeneration-specific candidate genes. Craig et al. examined the expression changes in regenerating photoreceptor cells, by combining light damage through continuous light of low intensity with lasercapturing of ONL tissue containing photoreceptors and their progenitors [55]. Samples collected at early time points after light damage showed decreased expression of photoreceptor-specific genes (such as the opsins) and up-regulation of stress response genes, indicative of dying photoreceptors. In a second step, transcript changes at 24– 48 h post-injury were examined, as they were thought to represent changes in progenitor cells migrating to the ONL. Differentially expressed genes found in this group included the transcription factors sox11b, fos, jun and pax6a as well as genes encoding secreted 367 growth factors such as l-plastin, CC chemokine, galectin family members, progranulin family members and midkines. In situ hybridization and immunohistological experiments of selected candidate genes showed that lgals9l1 (a member of the galectin family) and progranulin-a were exclusively expressed in microglia, and lgals1l2 in microglia as well as Müller glia cell, confirming the specificity of the approach. Whereas Craig et al. had isolated Muller glia cell-derived progenitor cells from the ONL, Qin et al. used the transgenic zebrafish reporter line Tg(gfap:GFP)mi2002 to directly isolate Müller glia cells [57]. Gene expression profiles from intact and light-lesioned retinas at 8, 16, 24 and 36 h post-treatment were compared. A set of 953 genes being differentially expressed was identified and subdivided into 3 clusters, depending on their temporal expression patterns. One set of transcripts was up-regulated immediately after the lesion, and included mainly genes implicated in translation and protein biosynthesis. A number of DNA replication and cell cycle genes were upregulated slightly delayed, reflecting re-entry of Müller cells into the cell cycle. Among the transcripts down-regulated after lesion were genes involved in chromatin assembly and ion homeostasis, consistent with dedifferentiation of Müller glia cells. As expected, an upregulation of cell cycle genes and growth factors such as pcna and pdgfa (platelet-derived growth factor) was observed as well. Interestingly, despite taking such a specific approach, Qin et al. identified two candidate genes, hspd2 (heat shock 60-kDa protein 1) and msp1 (monopolar spindle 1), which are not only essential for the regeneration of the retina, but likewise important for fin and heart tissue renewal in zebrafish [57]. Similar to hspd2 and msp1, the heparine-binding growth factors midkines a and b seem to play a significant role in regeneration of a variety of tissues in the zebrafish (retina, heart and fin) [54,58,59]. Photoreceptor apoptosis after light damage causes midkine expression pattern changes in the zebrafish retina [54]. In addition to their developmental expression (midkine-a in progenitors, Müller cells and horizontal cells and midkine-b in retinal ganglion and amacrine cells), both midkines are expressed in proliferating Müller glia cells and their progeny, and midkine-b is additionally found in HCs [54]. Other candidate genes identified in microarray experiments and hypothesized to be involved in the switch from differentiated Müller glia cell to proliferating retinal stem cell are stat3 (signal transducer and activator of transcription 3), ascl1a (ash1a) (achaete/scute Table 1 List of molecular cues involved in retina regeneration in the zebrafish. Gene Classification hspd1 (heat shock 60 kDa protein 1) mps1 (monopolar spindle 1) Mitochondrial protein chaperone involved in stress response Kinase involved in mitotic checkpoint regulation mdka (midkine-a) (secreted) heparin binding growth factor mdkb (midkine-b) (secreted) heparin binding growth factor ascl1a (ash1a) (achaete–scute Proneural gene homolog 1a) notch3 Component of Notch–Delta signalling deltaA Component of Notch–Delta signalling (notch ligand) pax6b Transcription factor (neurogenesis, regeneration) stat3 (signal transducer and Transcription factor activator of transcription) ngn1 (neurogenin 1) Transcription factor olig2 (oligodendrocyte bHLH transcription factor transcription factor 2) lgals9l1-, 2 (galectin family) Secreted growth factor progranulin-a Secreted growth factor Action Lesioning paradigm Reference Formation of retinal stem cells from dedifferentiating Müller cells Proliferation of photoreceptor progenitors Light damage [57] Light damage [57] Not clear yet Not clear yet Transition of Müller cell into proliferative stem cell Light damage Light damage Surgical excision [54] [54] [61] Progenitor proliferation Differentiation Surgical excision Surgical excision [61] [61] Hypothesized role in proliferation and migration of neuronal progenitor cells Hypothesized to trigger Müller cell proliferation in the damaged retina (has been shown to do so in the undamaged retina) Hypothesized role in “re-differentiation” of Müller cells Hypothesized role in maintaining Müller cells in a dedifferentiated state Up-regulated 24–48 h post-injury in microglia and Müller cells; exact function unknown Up-regulated 24–48 h post-injury in microglia; exact function unknown Light damage [37] Light damage [56,60] Light damage Light damage [37] [37] Light damage [55] Light damage [55] 368 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 homolog 1a) and members of the notch–delta family. Stat3 expression is increased in photoreceptors and Müller glia cells following light damage [56]. Interestingly, only a subset of Stat3+ Müller glia cells co-label with PCNA, which led Kassen et al. to hypothesize that Stat3 might act as a molecular cue, triggering MCs to proliferate after injury. In a recent publication, Kassen et al. identified a potential upstream regulator of Stat3, CNTF (ciliary neurotrophic factor). CNTF activates Stat3 to induce Müller cell proliferation in the undamaged retina [60]; whether or not it plays a similar role in the injured retina remains to be investigated. Yurco and Cameron found up-regulation of members of the notch– delta family, as well as ascl1a (ash1a) in their microarray analysis of differentially regulated genes after excision of a patch of adult zebrafish retina. Importantly, ascl1a (ash1a) was up-regulated before components of Notch–Delta signalling showed increased expression. They proposed a model in which Müller glia cells start to express ascl1a (ash1a) after injury and subsequently become divided into two sub-populations, a notch3-expressing population consisting of proliferating progenitors and a deltaA expressing population exiting the cell cycle and differentiating into neurons [61]. Morpholino knock-down of ascl1a (ash1a) inhibits Müller glia cell activation and consequently regeneration [62]. Although ngn1 (neurogenin-1) is expressed in proliferating progenitors and the transgene olig2:EGFP in dedifferentiated Müller glia cells [37], nothing is yet known about their role in retina regeneration. 2.1.3. Open questions The unique capability to regenerate all retinal cell types, combined with tractable genetics, makes zebrafish an ideal model for deciphering the molecular pathways underlying regeneration. Understanding zebrafish biology is valuable for developing future therapeutic interventions in human retinal disease. As evidence accumulates that the molecular components and stem cells implicated in retinal regeneration might be similar to those involved in regeneration of other tissues, retinal regeneration research will help us build our understanding of the general mechanisms of regenerative processes. Despite our progress a number of open questions remain. Differentiated Müller cells re-enter the cell cycle after injury to produce retinal progenitors; however, little is known about the extrinsic signals regulating lineage specification. In the undamaged retina, Müller glia cell-derived progenitors in the INL primarily give rise to rod photoreceptor progenitors; injury seems to induce a “fate switch” for Müller glia cells, turning them into progenitor cells for other retinal cell types. How this is achieved remains largely unknown. Another unresolved question is if all or only a subset of Müller glia cells are multipotent and can give rise to retinal progenitors. pax6, a gene associated with neurogenesis capacity and multipotency, is expressed at low levels in all differentiated Müller cells in the zebrafish retina; this may indicate that all Müller cells are able to dedifferentiate. Additionally, Fimbel et al. showed that the majority of MCs (90%) express PCNA at 3 dpi in the damaged retina [43]. Alternatively, there is compelling evidence indicating that it is only a subset of MCs switching to the progenitor fate. Kassen et al. showed that Stat3 is expressed only in a subset of Müller glia cells during regeneration [56]. Similarly, only a subset of BrdU+ cells in the INL colocalized with the Müller cell-specific marker CAZ upon light damage in the study conducted by Morris et al. [41]. Thummel et al. observed two distinct groups of MCs in the gfap:EGFP transgenic zebrafish line after damaging the zebrafish retina with intense light. While one group of EGFP+ MCs maintained normal Müller cell morphology and did not co-label with PCNA, another set of cells showed a reduced amount of EGFP expression and did co-label with PCNA. The second group likely represents MCs re-entering the cell cycle and dedifferentiating [42]. Although there is evidence that damage to photoreceptors and other cell types of the retina, trigger the regeneration response, the identity as well as the source of the actual “trigger signal” is not known. Fausett and Goldman speculated that the molecular signal arises from cells near the actual lesion site, as 90% of Müller glia cells next to the lesion site are GFP+, whereas at sites far from lesion, almost none are GFP+ [39]. Evidence in mouse models suggests that photoreceptors signal damage to Müller glia through endothelin receptors [63]. Finally, it is worth mentioning that Müller glia cells are not the only glial cell population in the zebrafish retina. Microglia are phagocytic cells that secrete pro- or contra-inflammatory signals to either inhibit or promote neuronal repair and regeneration [64]. Microglia hypertrophy at 24 h after light damage to photoreceptors has been shown. Subsequently, they become polymorphic and migrate into the ONL [16,55]. Craig et al. found members of the galectin and progranulin families up-regulated after light damage [55]. Progranulin-a, lgals1l1 as well as lgals1l2 are expressed in microglia thus suggesting that microglia likely carry out an important role in retina regeneration and warrant further study. 2.2. Optic nerve Injuries of the optic nerve have a highly detrimental effect on a patient's life, severely impacting vision and ultimately leading to complete blindness in a significant percentage of cases. Damage to the optic nerve can occur by either trauma (e.g. head trauma) or can be caused by diseases leading to death of retinal ganglion cells and subsequent axon degeneration, such as general neurodegenerative disorders (e.g. multiple sclerosis) or ocular disorders. The most prominent of the latter is glaucoma in which increased intraocular pressure is thought to lead to optic nerve damage. Glaucoma is the cause of 15–20% of vision loss worldwide and thus treatment and cure are major goals of public health research. Successful therapeutic interventions aim to halt nerve degeneration and ideally, trigger regeneration of retinal ganglion cells and most importantly, axon regeneration. However, mammals (including humans) are unable to regenerate an injured optic nerve under normal physiological circumstances. After injury, RGCs show only transient sprouting and no long-distance regeneration. Compiling the findings of a great number of studies in rodent models, we can conclude that failure of RGC axons to regenerate can be attributed to two major reasons: 1) Existence of a non-permissive/inhibitory environment and 2) Lack of growth-promoting molecules (in the environment or intrinsic to RGCs). Experimental therapeutic strategies typically target one of these two causes, either by trying to counteract inhibitors of axon re-growth or by activating the neuron-intrinsic capacity to regenerate (for a review see [65]). Studies in several species have shown that macrophage activation (e.g. by injuring the lens or using pro-inflammatory agents) can promote (partial) axon regeneration [66–69]. More recently, researchers are trying to apply combinatorial treatments. Using a gene therapy approach, Fischer et al. showed that activation of the RGC's intrinsic growth state while simultaneously overcoming inhibitory signals, significantly enhances axonal regeneration [70,71]. However, full regeneration has not been achieved in any rodent animal models and it is not clear if regenerated axons are able to correctly pathfind and innervate targets in a precise fashion. The development of novel therapeutic strategies certainly requires a thorough understanding of the molecular basis of optic nerve degeneration and regeneration (or the lack thereof). Clearly, alternative vertebrate animal models with an intrinsic capacity for regeneration will help us decipher how cell-intrinsic and micro-environmental inhibitory and permissive signals/factors drive V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 RGC axon regeneration. Interestingly, the molecular and cellular reactions to optic nerve injury observed in mammalian models seem to be inversed relative to fish in many aspects. In zebrafish we generally find an abundance of growth-promoting molecules. Inhibitory molecules, if expressed, usually exhibit a different pattern and/or timing of expression. This might partly be due to the fact that the eye continues to grow throughout the lifespan and thus the visual system is primed to continuously deal with newborn RGCs extending their axons to the optic tectum. It seems logical that signalling cues expressed during development will still be abundant in adulthood so they can be read by newborn RGCs. It has been clearly shown that after ON crush or lesioning, RGCs start to re-express molecules that they also express during embryonic development. These include cytoskeletal proteins, membrane-bound recognition molecules of the immunoglobulin family (e.g. NCAM, L1 homologs) and GAP-43 (growth-associated protein 43) (for a review see [72–74]). GAP-43 is localized to growth cones and presynaptic terminals of developing as well as regenerating axons [75]. Kaneda et al. examined the expression patterns of GAP-43 and its phosphorylated isoform (which is known to be active at growth cones) in the regenerating zebrafish optic nerve and observed a biphasic increase of phospho-GAP-43. Interestingly, the two distinct phases (the first one a steep increase in expression after 7–14 days, the second one a long plateau increase after 30–50 days) coincided with the recovery of two visuallymediated behaviors (OMR and chasing behavior) [76]. In the following section we will discuss recent studies aimed at achieving an understanding of the molecular basics of optic nerve regeneration in the zebrafish (see Table 2 for a list of genes). 2.2.1. Molecular signals/pathways implicated in optic nerve regeneration 2.2.1.1. Growth-promoting signals. Regenerating RGCs need two types of “positive cues” enabling them to extend their axons to their respective target regions: 1) molecular signals within the cell, switching (or maintaining) the cell to (in) a growth state, and 2) permissive and growth-promoting signals in the environment of the optic tract and tectum, allowing axon extension, pathfinding and correct target innervations. 2.2.1.1.1. Intrinsic growth-promoting factors. Whereas zebrafish continue to express a variety of genes needed for RGC development throughout their lifetime, mice often down-regulate their expression prenatally. In mice, several lines of evidence show that CNTF (ciliary neurotrophic factor) acts as an RGC-intrinsic axon growth factor and changes in CNTF expression levels during ON injury underlie RGC apoptosis. In untreated mice, CNTF receptor is lost from RGC axons after optic nerve crush and glial cells (astrocytes and oligodendrocygtes) start to express CNTF shortly after lesioning. It is thought that the loss of CNTF receptor on RGCs renders them insusceptible, leading to cell death, while the appearance of CNTF in glial cells promotes scarring [77]. These processes can be stopped by inducing intraocular inflammation, which induces retinal astrocytes to release CNTF, 369 switching RGCs to a regenerative state [78] or intravitreal injections of exogenous CNTF [79]. In cultured goldfish RGCs, CNTF in combination with two molecules secreted by optic nerve glia, axogenesis factor-1 (AF-1) and AF-2, drives axon regeneration [80]. In zebrafish, CNTF is known to play roles in photoreceptor neuroprotection and induction of Müller glia cell proliferation in the retina of undamaged zebrafish [60]. Its implication in ON regeneration however has not been studied thus far. Another class of proteins acting as RGC-intrinsic outgrowth molecules are the reggies (or flotillins). Morpholino-mediated combinatorial knock-down of the three zebrafish reggie orthologs 1a, 2a and 2b results in an overall reduction in RGC axon regeneration [81,82]. Similar to the situation in zebrafish, knocking down reggies in mouse hippocampal and N2a cells inhibits neuronal differentiation and neurite extension in vitro [81]. These findings indicate that this class of scaffolding proteins is essential for ON regeneration, most likely by acting as microdomains for the assembly of multiprotein signalling complexes recruited during regeneration. Members of the KLF (krüppel-like factor) family are clearly implicated in successful ON regeneration. They had been identified initially by microarray analysis of changes in gene expression during ON injury in zebrafish. Veldman et al. had used laser micro-dissection to isolate RGCs from uninjured and 3-day post-injury zebrafish and identified 347 up-regulated and 29 down-regulated genes. Out of six genes chosen for further investigations using morpholino-mediated knock-down, two out of those six, KLF6a and KLF7a, were necessary for robust RGC axon re-growth [83]. The same research group recently also identified KLF4 as a key player responsible for the loss of axon regrowth capacity in adult mice [84]. Mice with a tissue-specific loss of KLF4 in RGC cells were subjected to ON crush and assessed for regeneration after 2 weeks. Strikingly, knock-out mice showed an increase in the number of regenerated axons, however axons did not regenerate fully. The authors propose that different members of the KLF family act in concert to regulate the regenerative capacity of CNS neurons in mice. 2.2.1.1.2. Growth-promoting molecules secreted from a permissive environment. In addition to RGC-intrinsic molecular signals, cues from the ON microenvironment play a significant role in axon outgrowth. Glial cells are major players in axonal regeneration, as they secrete a variety of growth-promoting molecules. Examples are axogenesis factors 1 and 2, which have been shown to activate the expression of growth-related molecules via a purine-sensitive pathway in the goldfish [80]. Goldfish and zebrafish oligodendrocytes up-regulate the expression of members of the immunoglobulin family, such as L1, P0 and Contactin1a, after ON lesion [72,85,86], suggesting roles in axon regeneration and re-myelination. Interestingly, not only do glial cells seem to secrete permissive or inhibitory factors into the optic nerve microenvironment, but so do photoreceptors. Matsukawa et al. isolated purpurin as a candidate gene important in ON regeneration in goldfish by screening a cDNA library for genes up-regulated during regeneration. The site of purpurin expression was shown to be goldfish photoreceptors [87]. Table 2 Molecular cues implicated in successful optic nerve regeneration in the zebrafish. Gene Classification Action Reference reggie-1a, 2a, 2b Microdomain scaffolding protein (assembly of multiprotein complexes) Growth factor Transcription factors Growth factors Retinol-binding protein Axon guidance molecule Sulfated glycosaminoglycan Repellent axon guidance molecule Repellent axon guidance molecule Regulate axon regeneration, axon outgrowth and neuronal differentiation [81] Drives axon regeneration in combination with axogenesis factors 1 and 2 Thought to regulate axon regeneration Activate expression of growth-related molecules Induces neurite outgrowth Pathfinding Pathfinding Pathfinding Establishes retinotopy during regeneration [60,80] [83,84] [80] [87,88,109] [89] [100] [104,105,110] [106–108] CNTF (ciliary neurotrophic factor) KLF 6a, 7a (krüppel-like factor) axogenesis factor 1, 2 purpurin netrin-1 Chondroitin sulfates tenascin-R ephrin-A 370 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 Purpurin is a retinol-binding protein, and bound to retinol, induces neurite outgrowth in retinal explant cultures. This outgrowth can be inhibited by adding an RALDH2 (retinaldehyde dehydrogenase) inhibitor (disulfiram). This has led to the speculation that the action of purpurin might be mediated by retinoic acid synthesis. Nagashima et al. tested this hypothesis by looking at the expression levels of components of the retinoic acid synthesis pathway, such as RALDH2, CYP26a1, CRABPs (cellular retinoic acid binding protein) and RARs (retinoic acid receptor). RALDH2, CRABPs and RARS were significantly up-regulated after optic nerve injury in goldfish, whereas CYP26a1 (an enzyme degrading retinoic acid) was down-regulated to about 50% of normal levels. These results indicate that retinoic acid signalling might play a key role during the first stage of optic nerve regeneration in fish [88]. The axon guidance molecule Netrin-1 is promoting RGC axon regeneration in goldfish. Netrin-1 expression is maintained until adulthood and receptors are up-regulated on newborn RGC axons [89]. In contrast to zebrafish, its expression is developmentally downregulated in rats [89]. 2.2.1.2. Inhibitors. Although initial sprouting of neurites is observed in mice after optic nerve injury, RGCs die shortly thereafter through apoptosis, which is thought to be triggered by the activation of astrocytes initiating wound repair and scarring. Inhibitory signals can be roughly subdivided into myelin, glial scar and repellent axon guidance molecules. Myelin proteins are major inhibitors for axon regeneration in mammals [90]. Data on how myelin affects axonal re-growth in fish is somewhat unclear. Whereas the role of myelin has not been studied in zebrafish, findings from goldfish seem to argue that goldfish myelin is not a strong inhibitor [91–93]. The myelin-associated Nogo protein is a major inhibitor of axon re-growth in the optic nerve and spinal cord of mammals [94]. Knocking out Nogo A,B or C or inhibiting the Nogo receptor in mice abolishes this inhibitory effect and allows neurite extension into the ON lesion [95]. Although zebrafish do possess a Nogo A ortholog, the zebrafish isoform does lack the NogoA-specific N-terminal domain that inhibits regeneration in mammals [96]. A recent study examined the role of the second, C-terminal, inhibitory domain (Nogo-66) in zebrafish axon regeneration [97]. Binding of Nogo-66 to its receptor NgR has no inhibitory effect on zebrafish RGC axon growth and in fact is growth promoting. Another prominent source of inhibition in mammals is the glial scar formed in response to optic never injury. A hallmark of the glial scar is increased immunoreactivity for chondroitin sulfates, which has been shown to inhibit axonal regeneration [98]. In an elegant experiment, Charalambous et al. grafted chick neural tube stem cells onto a lesioned rat optic nerve, and observed the induction of MMP (matrix metalloproteinase)-2 and -14 in astrocytes of the optic nerve. This in turn led to the degradation of matrix chondroitin sulfate proteoglycans and allowed rat axons to migrate towards the thalamus and mid-brain [99]. Chondroitin sulfates are also present in the developing and adult non-retinorecipient pretectal brain nuclei of zebrafish. However, they are not found before or after an optic nerve crush in the optic nerve and tract in adult zebrafish [100]. Quite different to the situation in rodents, where chondroitin sulfates seem to play a direct role in inhibition of axon regeneration, Becker and Becker postulate that in zebrafish, chondroitin sulfates play a role in axon regeneration by acting as repellent axon guidance molecules that help axons finding their correct path to their tectal targets [100]. As the zebrafish retina continues to grow throughout the life of the fish, newborn RGCs are constantly added to the retina and their axons need to find the correct path along the optic nerve. Developmentally regulated axon guidance molecules thus might remain in the ON microenvironment and play a key role in regeneration. Semaphorin3a, a major repellent axon guidance cue, is upregulated after axotomy of the rat ON for 28 days [101]. Calander et al. reported the expression of Sema3Aa in the retina, optic chiasm and optic tectum in zebrafish. However, it does not seem to be present during the time window when the axons extend [102]. Becker and Becker state that neither of the two zebrafish 3A orthologs are detectably expressed in the lesioned optic nerve tract in the adult zebrafish (unpublished results reported in [73]). The lack of expression of this repellent axon guidance molecule in zebrafish might explain why axons are regenerated. While the lack of repellent axon guidance molecules in zebrafish might help newly generated axons to extend, the expression of repellent guidance cues in specific patterns has been shown to be essential for establishing correct pathfinding and retinotopy. TenascinR is an extracellular matrix molecule acting as a repellent guidance cue during development [103]. Areas of TenascinR expression are also maintained in the adult zebrafish, bordering newly growing axons from the retinal periphery [104], indicating that it also is implicated in guiding regenerating axons to their respective brain targets. In mice, TenascinR similarly seems to act as a repellent guidance molecule. Outgrowth of embryonic and adult retinal ganglion cell axons from mouse retinal explants is significantly reduced on homogeneous substrates of TenascinR [105]. Becker et al. have observed an increase in the number of cells expressing TenascinR mRNA in the lesioned optic nerve and have hypothesized that the continued expression of the protein after an optic nerve crush may contribute to the failure of adult retinal ganglion cells to regenerate their axons in vivo. Another class of repellent cues involved in establishing retinotopy in the tectum are the Ephrin-As, acting via Eph receptors. Developmental gradients of Ephrin-A2 and -A5 are retained in unlesioned and lesioned adult zebrafish [106], and blocking Eph receptors in goldfish has a detrimental effect on establishing retinotopy during regeneration [107]. Interestingly, while EphA3 and A5 (receptors) have been found to be re-expressed in a gradient in the retina of goldfish [108], their expression is not recapitulated during optic nerve regeneration in zebrafish [106]. 2.2.2. Open questions Data about molecular factors influencing optic nerve regeneration are accumulating and provide starting points for new therapeutic approaches. Recent gene therapy trials conducted in mouse models have shown that combinatorial treatments can switch RGCs to a growth state and lead to some extent of axon regeneration [111–113]. Despite this progress, several open questions remain: Are the newly regenerated RGC axons going to innervate the appropriate pretectal and tectal targets and re-establish a correct retinotopic map? Moreover, ultimately, will regeneration of RGC axons lead to recovery of visual function? Strikingly, the answer seems to be “yes” in zebrafish and goldfish. RGC axons not only re-grow, but also find their way back to their original targets, leading to full recovery of visual function [72,114– 117]. Unfortunately, not much is known about the molecular pathways underlying restoration of vision, as research performed to date has mainly focussed on the first steps in ON regeneration, axon regrowth and pathfinding. Integration of regenerated axons into existing circuits is a prerequisite for restoration of vision in human patients. Therefore, a thorough understanding of these late-step processes in regeneration on a molecular level is indispensable for successful future therapeutic approaches. The zebrafish model system has great potential for uncovering the molecular pathways underlying network integration of regenerated optic nerve fibers. 2.3. Spinal cord Abundant research efforts to facilitate spinal cord regeneration in a clinical setting have focussed largely on rodent models, providing valuable insights into the cellular processes required for functional recovery. Despite intense efforts, the limited ability of mammalian V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 spinal cord neurons to generate new cells and form new connections remains a limiting factor. Thus the goals of basic research, to allow translation into the clinical setting, include the enhancement of regeneration by identifying interactions that alleviate these restrictions or activate latent cells. Anamniotic vertebrates exhibit substantial spinal cord regenerative capacity and can thus provide insight into the events that underpin successful functional regeneration. Recent reviews and volumes compare the utility and contributions of anamniotic vertebrates to spinal cord regeneration [118,119]. Clinical sources of spinal cord injury typically include trauma and neurodegenerative diseases. Principal issues in clinical spinal cord regeneration include: i) protection of surviving cells from secondary degeneration following injury; ii) proliferation of new cells; iii) growth of axons through/around scar; iv) connection of axons to original targets; and v) plasticity of the CNS to restore function despite imperfect pathfinding. Spinal cord regeneration from stem cells will require multimodal treatment to address neuron-intrinsic and neuron-extrinsic limitations. Following spinal cord injury in mammals, the number of new axons generated is very limited. In sum, it appears that the plasticity of spared axons is the principal substrate for the limited functional recovery that is achieved [120]. Thus, studies of enhanced and directed axonal sprouting will be valuable to implementing treatments. Regeneration of neurons will be especially important to human corticospinal lesions and restoration of voluntary motor control, because neuronal sprouting is limited in these areas [120]. An understanding of neuronal differentiation and functional integration is of high priority and a number of ongoing clinical trials using stem cell therapy seek to address these issues [121–123]. Zebrafish, with their capacity to regenerate neurons and their axons and regain function following experimentally induced spinal cord injury, offer promise in addressing several key questions. Beyond the various advantages of the model system, studying functional regeneration of the spinal cord in zebrafish is expected to be especially fruitful. Drug and gene discovery in regard to spinal cord regeneration are very practical in zebrafish because measurements of functional recovery can be automated in this highly active species (reviewed below). Unfortunately, their small size will not permit analysis of axon pathfinding to navigate large injuries. In the context of spinal injuries observed clinically, larger rat models may be considered too small. The size of an adult zebrafish is an order of magnitude less than the size of a typical human spinal cord lesion. However the small size of the organism is an advantage in many regards, as in most model organisms, providing economy and generation time efficiency not practicable in larger animals. In sum, it appears that zebrafish will be invaluable to discovering targets for functional regeneration of the spinal cord, serving as a very effective compliment to murine and large-animal models. Studies of the zebrafish spinal cord have demonstrated robust regenerative capacity. Detailed research differentiates cell and injury types that influence the regenerative response and delineates the molecules that underpin such differences. Regeneration of zebrafish adult spinal cord neurons from most of the spinal-projecting brain nuclei is observed following spinal cord lesion [124]. Various nuclei display different regenerative properties, including the molecular pathways engaged, depending on the rostral–caudal position of the lesion [125,126] and identifiable patterns within this heterogeneity may be exploited to understand molecular pathways determining the intrinsic ability to regenerate. Pathways underpinning regeneration seem to be primarily those previously noted for similar functions in mammals and almost exclusively focus upon cell-adhesion proteins. The immunoglobulin superfamily cell-adhesion protein L1.1 contributes to spinal neuron growth, synapse formation and functional recovery in zebrafish [127], as demonstrated by applying a morpholino directed against L1.1 to the spinal cord lesion. This effective gene knock-down technology is an economical and informative loss-of- 371 function complement to tests in mammals, where L1 has been shown to be up-regulated during regeneration in CNS nerve grafts and L1 fusion protein application can promote functional recovery of locomotion [128,129]. Similarly, GAP-43 is up-regulated in zebrafish spinal cord regeneration [125], as in mammals [130–132]. On the other hand, another cell-adhesion molecule, protein zero (P0, a component of the myelin sheath) shows substantive difference between anamniotes with regenerative capacity when compared to mammals with limited spinal cord regenerative potential [85,133]: This and difference to mammals allow studies in zebrafish to contribute to understanding re-myelination following regeneration [133]. Recent work has examined another cell-adhesion molecule, Contactin1a, which is up-regulated in both neurons and oligodendrocytes during spinal cord lesion [86] and is discussed above in the section on optic nerve regeneration. Although many classes of neurons of the zebrafish regenerate following spinal cord lesion, some do not. One example of cells with limited ability to regenerate their axons is the Mauthner cell [118,124]; a large and easily identifiable cell type that is amenable to electrophysiological recording in larval zebrafish and controls swimming [134–137]. Because other neurons consistently regenerate via the glial scar following lesion, Bhatt et al. have suggested that the Mauthner cell has intrinsic properties (a surface protein or a component of its intracellular signalling cascade) that limit its regenerative capacity [138]. It is noteworthy that if such an intrinsic factor exists then it is presumably limiting regeneration through detection of some property in the neuron's environment. If this is the case, identification of molecules extrinsic to the Mauthner cell environment that interact with Mauthner-specific sensors is critically important. Although Mauthner cell regeneration is rare, it can be bolstered by altering intracellular signalling, including application of cAMP [138]. cAMP is known to affect neuronal sprouting and regeneration in other systems although which of the many pathways it may be modulating remains unclear. Mauthner cell regeneration is an attractive target from a technical point-of-view, as functional regeneration (reviewed below) can be assessed via trunk movement [138] related to escape behavior, with the potential for high-throughput screening in 96-well plates [139–141]. These large cells can be recorded from or stimulated readily for electrophysiology [138]. Heterogeneous glial reactions that are expected to affect regeneration of spinal cord axons in mammals also occur in zebrafish [125]. Macrophages/microglia increase in numbers and phagocytose myelin debris, which may or may not be beneficial to axon regeneration [142]. In mammals, such differences have been identified to underpin the permissive or restrictive properties of glia towards regeneration [143]. 2.4. Brain nuclei Clinical regeneration of neurons in higher brain centres is a laudable, oft-described goal of stem cell biology. Ambitious targets of stem cell treatment include neurodegenerative diseases including Alzheimer Disease, Parkinson, and ALS [144–147], as well as stroke and traumatic brain injury. Indeed, proliferating cells may not only be expected to replace lost neurons, but may also be able to support the survival of remaining cells. Zebrafish are attracting considerable interest in understanding cellular and molecular pathways directing radial glia to become stem cells, and directing CNS stem cells to replace lost neurons. Adult neurogenesis is substantial in the zebrafish brain [148–156] and includes brain centres beyond the restricted forebrain proliferation typically studied in mammalian models. Compared to mammals, zebrafish adult neurogenesis is located in multiple centres, but is fundamentally similar in process; subventricular proliferative zones give rise to nascent neurons that migrate radially to their differentiated 372 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 destination. There is a notable lack of published experimental paradigms in which zebrafish brains are damaged, thus allowing the regenerative response to be addressed. Nevertheless, these approaches are certainly on the horizon [157], building on the substantial research in other fish models [158]. Efforts to define zebrafish adult stem cell properties will be very valuable in studies of functional regeneration. Stem cells of the adult zebrafish brain have recently begun to be defined, establishing zebrafish as a valuable model for adult brain stem cell biology. In the tegmentum, her5 (orthologue of mammalian Hes) cells meet the definitions of bona fide stem cells (including selfrenewal, slow proliferation and multipotency) and H/E(Spl) factors generally have a role in maintaining a neural progenitor state. These populations express markers of stem cells conserved with mammalian neural progenitors, including BLBP, GFAP, and Sox2. Presumptive neurons arising also express conserved markers, including asha1, her4, ngn1, and Delta [159,160]. 3. Lesioning paradigms Animal models of neuronal regeneration necessarily depend upon the manner in which the neurons (and neighbouring cells) are damaged. Various paradigms have been developed in an opportunistic fashion, each with different goals and suite of advantages. In some instances, experimentally induced damage may have a central goal of trying to replicate a disease process. Although a plethora of animal models of disease exist, few zebrafish models of disease have been brought to bear on regeneration. Instead, most work have focussed on an experimental intervention (surgery, neurotoxin application, etc., reviewed below) that strives for reproducibility. Traditionally such studies have been opportunistic in that they ablate accessible cells. For example the optic nerve and spinal cord are accessible to surgical lesion, and the retina is accessible to neurotoxin delivery via intraocular injection, or to toxic light lesion in albino animals. We review here methods that have been informative for regeneration studies in the past, and look ahead to transgenic paradigms that are well suited to the zebrafish model and have the potential to allow cell ablation/regeneration to be studied within a vast variety of neuronal cell types. We end with a commentary that assessing the function of regenerated neurons needs to become the benchmark of the field. Thus physiological and behavioral measures of regeneration ought to be a substantial component of assessing regeneration, and thus in the design of CNS regeneration paradigms. 3.1. Advantages to refining methods A variety of lesioning paradigms have been employed for the induction of neuronal cell ablation with goals of learning more about regeneration of the zebrafish CNS. Techniques range in the specificity of lesioning (Fig. 2). Whereas surgical excisions of pieces of tissues are not cell-specific, methods such as laser or light lesioning are potent to target specific layers of the CNS. These techniques have allowed research groups to observe the regeneration process, but a pitfall is that many cell types are ablated and thus questions concerning specific cell fate determination cannot be readily answered [23,24,161]. Cell ablation using a laser can also be used for targeting only specific neuronal cells of interest in a tissue [162,163]. The use of chemical toxins, such as ouabain, injected into the eye also possesses a certain level for controlled cell death. The concentration of the injected toxin dictates the extent of penetration of the toxin into the retinal layers, and therefore the amount of cell death [43,164]. Novel Fig. 2. Multiple methods of neuronal ablation to study CNS regeneration in zebrafish. The method of cell ablation influences the complexity of regeneration and thus experimental design. The specificity of the neurons targeted (ablation of only one or some neuronal cell types versus ablation of an entire section of retinal tissue) is dependent on the technique: A) light/laser lesion, B) chemical ablation, C) surgical lesion or D) NTR transgenic ablation (see also Fig. 3). V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 techniques employing transgenic-chemical ablation of neurons are proving to be the most specific means for targeted cell ablation [165]. Regeneration paradigms can be refined through the standardization of the lesioning paradigms. Studies will become more comparable between treatments or between research groups when protocols are performed with exacting rigor, eliminating some of the variables such as the unwanted death of neighbouring neurons or the activation of different cell death signalling pathways (i.e. apoptosis versus necrosis) due to distinctive lesioning methods, both of which could have an impact on the regeneration of the neurons of interest [166,167]. Refinement of techniques also leads to simplifying the ability to visualize cell ablation and regeneration in vivo [165]. With each investigation of regeneration, the possibility of applying this knowledge towards therapeutics for the treatment of human neurodegenerative diseases gets closer to becoming a reality. Characterization of the biological mechanisms and signalling pathways involved in this process is crucial. Thus refined strategies for neuron damage/ablation, each with its particular advantages, are a worthwhile consideration. 373 paradigms in other tissues (e.g. surgical intervention in heart and tail fins). In studies of regeneration of the vertebrate spinal cord, researchers have attempted to simulate human spinal cord injury in an effort to discover potential therapies. Many investigations have employed teleost fish including adult zebrafish, cichlids and goldfish, and similar surgical techniques have been used to lesion their spinal cords [124,170,179]. The vertebral column is exposed with a longitudinal incision, then is transected in the region of interest [124]. A similar technique to study spinal cord regeneration in mammals has been performed using adult rats and mice [169,172], and regeneration studies of Xenopus tadpoles have performed tail amputations to allow for comparisons of tail regeneration with spinal cord regeneration [173]. In an investigation by Becker et al., axonal regeneration of the zebrafish spinal cord was studied. After transection of the cord, regrowth of projections from brain nuclei can be observed. Within six weeks post lesion, most of the brain nuclei had axons that extended past the transaction and into the distal spinal cord, indicative of regeneration [118]. 3.2. Methods of inducing neuronal death and damage 3.2.1. Surgical lesioning Surgical lesioning of CNS tissues has been performed on the optic nerve, spinal cord, brain and retinal neurons of many organisms including teleost fish, mouse, rat and Xenopus [33,53,77,124,158,166–174]. Optic nerve crush is a surgical lesioning approach that is routinely applied in studies using rat, mouse and teleost fish. The most common surgical technique is to expose the optic nerve by gently pulling the eyeball out of the socket then crushing it, taking care not to sever the ophthalmic artery [77,100,106,166,167]. Optic nerve damage can easily be visualized and confirmed. In zebrafish, the optic nerve usually has a whitish opaque colour, but upon damage, a translucent stripe across the optic nerve becomes visible where the forceps had made contact [106]. Becker and Becker examined regeneration of the optic projection in zebrafish after optic nerve crush to study the role of genes involved in axonal growth, pathway selection, and target recognition [106]. The axonal projection is retinotopic, meaning that it is formed from/ composed of retinal ganglion cells projecting to the optic tectum [175]. After lesioning, certain tectal areas were observed to be primary targets for the regenerating axons in that there was an increased probability for functional integration of regenerated tissue. Becker and Becker found that the pattern of expression of the recognition molecules ephrin-A2 and ephrin-A5b correlated to the target areas, suggesting that these molecules may play important roles in functional regeneration of the optic nerve [106]. Although a major contributor to our understanding of nerve regeneration, the relevance of optic nerve crush to human disease is questionable. A significant advantage of this paradigm, however, includes the ability to trace pathfinding and target accuracy of the regenerated axons (reviewed above). Retinal regeneration has been investigated by surgically removing portions of the retina or the retina in its entirety [176,177]. Mensinger and Powers investigated the effects of removing parts or the entire retina on regeneration in the teleost models goldfish and sunfish [178]. While excision of the whole retina completely prevented regeneration, leaving only 5% of retinal tissue intact led to restoration of functional vision, as assayed by ERGs and dorsal light response (DLR). Surgical lesioning of the retina has also been used as a means to study gene expression profile changes during retinal regeneration [53]. Retinal damage via surgery has been useful for observing neuronal regeneration, but is less powerful for studying only a specific subset of retinal neurons because all neurons are regenerating at once [177]. It is noteworthy that this type of lesion leads to formation of a blastema during regeneration, and in this aspect may be comparable to lesion 3.2.2. Laser lesioning Laser lesioning has been applied to induce ablation of neural cells in various tissues and organisms [23,161,180]. Regeneration of the retina of teleost fish has been investigated by ablating photoreceptors with an argon laser. The argon laser promotes accurate placement of the lesions and can be used to ablate photoreceptor cells in the outer nuclear layer in the retina. In this procedure, the lens of the fish is surgically removed in order to gain access to the retina for focussing the laser. The resulting lesions appear as bleached spots on the retina [23,161]. A particularly advantageous feature of laser lesioning is that it can be employed on a cell-specific level to allow for regeneration studies of subgroups of neurons. Moreover, it can be directed to ablation of any part of the CNS. The targeted cells are labelled with an indicator such as a fluorescent dye to allow for easy visualization of neurons. In their study of the zebrafish hindbrain, Liu and Fetcho labelled cells through retrograde uptake of the indicator CGD by injecting it into the spinal cord of 4dpf fish [163]. After intense light exposure, the dye will have adverse affects, and thus cause cell death. The targeted neurons were identified by position and morphology, and subsequently injured by exposing them to the maximum intensity of the laser. Immediately after light treatment, fluorescence was observably decreased, and by 24 h after treatment it was no longer detectable in confocal images, suggesting cell death [163]. A similar approach was used on optic tectum in zebrafish larvae, utilizing effective measures of visual function [181]. Moreover, Kimmel et al. utilized laser irradiation on trigeminal neurons to investigate if their input to the hindbrain Mauthner cell affects its dendrite outgrowth [182]. These laser lesioning paradigms have not yet been used for regeneration studies, but appear to be an attractive method for this purpose. Finally, it is noteworthy that femtosecond laser pulses have been employed to ablate specific cells in live zebrafish [183]. Although this has not yet been used to ablate neuronal cells of zebrafish, it has been successful in studying neuronal regeneration in Caenorhabditis elegans [184]. Developing this technology, and combining it with the optical transparency and transgenesis of zebrafish have substantial potential for specific ablation and neuronal regeneration. 3.2.3. Light lesioning Light-induced lesioning is a popular method of cell ablation because it targets photoreceptor cells while leaving the remainder of the retina relatively unharmed, with less inflammatory response compared to tissue injury [185]. This system has been shown to be very effective for studies of the teleost retina, and is the most 374 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 commonly used lesioning paradigm for assessing regeneration in both albino and pigmented fish [186–188]. Mechanisms of cell death in light lesion have been reviewed extensively elsewhere [187,189] and likely include both generation of reactive oxygen species and accumulation of heat leading to photoreceptor death. Fish are initially dark adapted for a particular period of time, followed by the application of constant light, which induces photoreceptor cell death through apoptosis [24,40,185,190]. The extent of retinal injury to the organism is affected by the fish's history of light exposure, the period, quality and intensity of light exposure, and the organism's pigmentation and genetic background [24,187,191]. Regeneration of photoreceptors following such lesions is substantial (please refer to the Retina section of this review). 3.2.4. Chemical ablation Another technique used primarily for the study of teleost retinal regeneration is to expose the tissue of interest to toxic chemicals. Ouabain, a metabolic poison inhibiting the Na+/K+-ATPase, has specifically been employed to induce degeneration of cells in the retina [43,164,192]. Depolarization of the plasma membrane triggers a cascade of enzymatic events, leading to cell death [193]. The extent of cell death is proportional to the concentration of intravitreal toxin [43,164,194]. High concentrations of ouabain have been shown to rapidly cause death to all retinal nuclear layers, while the use of lower concentrations can be useful in biasing ablation towards the inner nuclear layers while leaving the photoreceptor layer relatively unscathed. Fimbel et al. took advantage of this element and demonstrated that selective damage to cells of the INL by injection of low doses of ouabain could stimulate regeneration [43]. A zebrafish model for whole retina regeneration has been presented by Sherpa et al. who showed regeneration of RGCs after cytotoxin injection [194]. Intravitreal injection of 6-hydroxydopamine (6-OHDA) has been used to ablate retinal dopaminergic cells in both goldfish and zebrafish [195–199]. Almost 3 decades ago Negishi et al. had shown that loss of dopaminergic neurons in goldfish retinas triggered increased proliferation of retinal progenitors [198]. Li et al. have used a behavioral test (escape resonse) to study the effect of loss of dopaminergic neurons on visual sensitivity in the zebrafish [195]. 3.2.5. Transgenic expression of toxic proteins Ablation of particular cells has been achieved by expression of disease genes or other toxic proteins in subsets of zebrafish photoreceptors [200,201]. This genetic ablation of cones versus rods has led to important insights into retinal progenitor pools in the zebrafish retina [202]. In contrast to emerging transgenic methods reviewed below, in which the toxicity is inducible and restricted to the timeline of prodrug exposure, the continued fruitful nature of these methods will be limited somewhat by the toxic protein being abundant in the regenerated cell. Arguably, the early steps of cell regeneration will be unperturbed in these models, and late-stage molecular signatures and functional measures of regeneration will need to consider whether the results are influenced by the toxic proteins used for ablation. 3.2.6. Transgenic conditional cell-specific ablation Genetic engineering techniques have become very practical in the zebrafish, and with the advancement of these tools, mechanisms of development and regeneration have become readily observable. Recently, investigators have created conditional cell ablation techniques by combining genetic and chemical tools to target the degeneration of particular cell types (Fig. 3) [165,203,204]. By specifically directing ablation of a particular neuronal cell class for a limited period of time, rather than causing simultaneous death to many types of neurons, as had been the standard procedure thus far, alterations and disruptions in relevant genetic pathways can be analyzed for their potential roles in CNS regeneration. One recently developed method of transgenic cell-specific ablation is nitroreductase-mediated conditional cell ablation [165]. Transgenic zebrafish express the Escherichia coli gene nfsB encoding the nitroreductase (NTR) enzyme. NTR is driven by a cell-specific promoter, resulting in its expression in the targeted neuron type. These neurons will function normally until the application of a prodrug, metronidazole (MTZ). MTZ binds to NTR and is electrochemically reduced, converting it into a DNA cross-linking cytotoxin [165] (Fig. 3B). The toxic form of MTZ remains isolated in the neurons of interest and induces base transitions, transversion mutations, and finally fragmentation of DNA [205]. This results in precise cell ablation of the subset of neurons expressing NTR, while other neighbouring cells appear to be unaffected by the toxin [165]. Ablation is discontinued upon the removal of the prodrug, allowing for regeneration to occur devoid of the influence of the toxin. The method has recently been used to ablate specific classes of bipolar cells in the zebrafish retina [204]. Based upon cell counts, the ablated bipolar cells reappear in the retina, presumably regenerating from a proliferating population. Another study has utilized a similar approach to ablate rod photoreceptors and demonstrated a differential proliferative response based on the abundance of rods ablated [241]. This study underlines the potential of the system, allowing specific cell classes to regenerate among the complexity of cell types in the retina. Certain questions remain to be tested, including whether adjacent cells are damaged or killed by the method, and whether the cell types regenerated are more or less diverse than those that are ablated. The specificity of the system is tantalizing, as it ought to allow a simpler system to define molecular mechanisms during regeneration of a single cell type rather than many cell types. Ideally, via comparison of regeneration of multiple classes, the community will learn triggers of cell-type-specific regeneration and functional rewiring. The methods used to date in zebrafish using nitroreductase cell ablation also utilize the Gal4-VP16 and UAS combinatorial transgenic system [165,204]. This allows the exciting potential for combination with Gal4 enhancer trap lines [206–208] with many neuronal and glial types now available for ablation in a variety of CNS tissues. Although the toxicity and transcriptional effects of VP16 itself need to be considered, this model system has enormous potential to efficiently ablate single or multiple cell types to compare their regeneration and functional integration. 3.3. Regeneration has limited utility if function is not restored The regeneration of a neuronal cell is an impressive process, but would be of limited utility if the cell does not functionally integrate into the existing tissue system upon regeneration. This aspect is often overlooked in regeneration studies. Neuronal regeneration in the zebrafish can easily be observed, but the functionality of the regenerated cells is often assumed but not experimentally tested. Defining the molecular mechanisms that specify the regenerating cell's fate will be an ongoing focus of the community, and will certainly contribute to clinical specification of stem cells. On the other hand, specifying cell fates alone will not be useful if the cells are not directed to reconnect into the neuronal network. Here we review regeneration paradigms that can be coupled with informative and/or efficient methods to assess the regenerated function using electrophysiology and/or behavior. 3.3.1. Electrophysiology output The ERG (electroretinogram) is a routinely used and extremely valuable tool for studying degeneration and dysfunction of retinal neurons in the zebrafish. This technique measures light-induced changes of electrical activity in the eye [11,209,210]. The ERG is a method for probing outer retinal function and can be adapted to larval zebrafish and adult zebrafish, as well as multiple other organisms. In V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 375 this non-invasive procedure, a recording electrode is placed onto the cornea of the zebrafish, and a response is induced by flashing light at specific intensities and durations [11]. The observed response is recorded by the instrumentation and interpreted into an ERG trace (Fig. 4) that has a form very simlar to the ERG of mammals. The vertebrate ERG output is expressed as three distinct waves (Fig. 4). An initial low-amplitude negative a-wave reflecting photoreceptor activity is followed by the large positive b-wave, mainly representing second-order (ON bipolar) neuron activity. The d-wave appears at light offset and is indicative of postsynaptic activity involved in the OFF response [11,211]. The ERG has been used in goldfish to assess functional recovery of retinal neurons after cytotoxic lesioning and surgical extirpation of the adult retina [178]. Recovery of the ERG a-wave up to 90% of control levels coincided perfectly with the regeneration of photoreceptors, and more importantly, also the b-wave amplitude recovered up to 50% in amplitude, demonstrating that regenerated photoreceptors made proper synaptic connections with second-order neurons [212]. The ERG has also been used to demonstrate functional integration during regeneration of a specific cone class in rainbow trout [191]. Restoration of function has also been established beyond the retina during cone photoreceptor regeneration in trout [213]. These recordings from the optic nerve where it enters the tectum have the advantage of demonstrating connectivity to higher brain centres, but require surgery and thus preclude documenting regeneration within the same individual. Electropysiological tools can also be employed for detecting electrical activity in the spinal cord of zebrafish. Patch-clamp electrodes are placed extracellularly on the zebrafish in order to record the locomotive response pattern [134,214]. 3.3.2. Behavioral output Behavioral responses are another mode to assess functional integration of regenerated neurons into existing circuits. The loss of certain behaviors upon the ablation of neurons, as well as the recovery of those behaviors once the cells have regenerated, can be assessed by a number of well-established paradigms in the zebrafish. Typically, these have focussed on innate responses of fish to presented stimuli, rather than operant conditioning, thus simplifying the methods considerably. Indeed the simplicity of many of these assays should make them accessible to most any research group. 3.3.2.1. Regeneration of motor behaviors. Regeneration after spinal cord transection in adult zebrafish has been studied employing a simple swimming performance assay [215]. 1 and 3 months after lesioning, fish were transferred into a tunnel with water flowing at slow and fast speed to test for swimming endurance as a function of fiber re-growth [215]. Other examples of robust assays include the light-induced startle response used in high-throughput screening for motor deficits [139– 141]. This assays certainly could be used to assess replacement of functional connections during hindbrain or spinal cord regeneration. Escape responses in zebrafish larvae can also be induced with a light touch or puff of water directed at the trunk. This type of assay has Fig. 3. Regeneration of a single neuronal class can be achieved following conditional cell-specific ablation. To reduce the heterogeneity of the regenerative resonse, one can reduce the complexity of the cell types that are ablated. This has recently been achieved using neuron-specific transgenic expression of the bacterial gene nitroreductase (nfsB, produces NTR protein), which kills cells upon application of the prodrug metronidazole (MTZ) [204]. A tissue-specific promoter fused to green fluorescent protein (GFP) drives expression of NTR in bipolar neurons of the retina (A). Cells expressing NTR develop normally until MTZ prodrug application, e.g. in a bath treatment (B). Upon application, MTZ is converted to toxin only in cells expressing NTR, leading to cross-linking of DNA and cell death (C,D). Removal of MTZ permits cell-specific regeneration, including expression of the fluorescent marker (E). This method can be combined efficiently with enhancer traps, allowing regeneration to be studied in a variety of cell populations (see text). 376 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 Fig. 4. CNS regeneration studies should emphasize functional integration of the regenerated cells. The electroretinogram (ERG) is an example of a non-invasive technique that can be applied to measure neuronal function, thus allowing withinindividual examination of the regenerative process (e.g. measures before, during, and after cell ablation). Analysis of the ERG response typically focusses on the amplitude of the b-wave (bold, blue positive deflection) in response to light stimuli. The b-wave primarily represents bipolar cell activity, and thus regeneration of this response can be diagnostic for appropriate integration of photoreceptors and bipolar cells following regeneration [191]. been used to assess functional regeneration of the Mauthner cell in larval spinal cord [138]. 3.3.2.2. Visually-evoked behaviors. Visually-evoked behaviors of zebrafish (and goldfish) can be easily stimulated and are readily accessible for measurement and genetic analysis of functional vision [12,216]. The optokinetic response (OKR) is a robust assay for visual function measuring zebrafish eye movements in response to a moving stimulus (grating) [217–220]. Ablation of cone photoreceptors for example, would be expected to impair colour detection in zebrafish larvae. Consequently they would not be able to track the movements of coloured objects. Cone photoreceptors' regeneration would lead to recovery of functional colour vision and improved performance in the OKR assay could be observed. The OKR has been used to study retinal regeneration after ouabain injection in the goldfish [221]. Cytotoxin injection resulted in lack or greatly reduced OKR performance. Regeneration of retinal cell layers resulted in partial recovery of several visually-evoked assays (ERG, DRL and OKR), however visual sensitivity did not completely return to control levels [221]. In the optomotor response (OMR), motion in a particular direction is simulated and zebrafish will swim in the direction of the perceived motion [222,223]. Failure of zebrafish to respond to the OMR stimulus can be attributed to distinct/different underlying causes, and thus the OMR can be used for studying regeneration of various CNS tissues. Swimming arrest will occur following lesion, and functional regeneration should be able to (at least partially) restore the optomotor response. As the OMR is dependent not only on a functional retina, but also on the correct propagation of visual stimuli to the fish optic tectum, regeneration of the optic nerve can also be tested. Kaneda et al. observed that the expression increase of the growth cone marker GAP-43 after optic nerve injury temporally concurred with the recovery of two visually-mediated behaviors (OMR and chasing behavior) [76]. 4. Regeneration as a potential confound when studying neurodegeneration International efforts are focussing on developing zebrafish as models of neurodegenerative disease, by identifying zebrafish mutants and deploying human versions of disease-associated genes (see other papers in this volume). These efforts parallel those in mouse and rat models, which have contributed much to understanding of disease progression [224]. It is noteworthy that such models, where neuron death is a benchmark outcome, are anticipated to exhibit substantial regenerative response. Regeneration from intrinsic stem cells during degeneration will represent both confounds and opportunities to modelling neurodegenerative disease. In particular, longitudinal studies assessing cell loss by quantifying neuron populations may be difficult if neurons are replaced at a rate similar to cell death. Perhaps such regeneration underpins the lack of obvious neuronal loss, despite neuropathological change in teleost models of light-induced photoreceptor death [225] or prion infectivity [226] (and other manuscripts in this volume). Instead, emphasis may be best placed on methods that assay cell death via expression of available cell death markers. Gene discovery approaches using zebrafish disease models may also need to consider regeneration in the way experiments are designed. For example proteomic, transcriptomic or metabolomic studies of disease models will need to interpret changes in molecular abundance in light of both degeneration and regeneration. One strategy to address such issues may be to focus experiments on aged fish, e.g. zebrafish more than two years old, where somatic growth and adult neurogenesis have slowed and the regenerative response is expected to be less robust. Beyond experimental design, other fundamental processes in neurodegeneration may be different in an organism that is capable of replacing lost neurons. These may include the cellular and molecular aspects of astrocyte reactions to injury. As reviewed above, populations of glia dedifferentiate to a progenitor cell state following zebrafish neuronal injury. This is in marked contrast to mammalian systems and thus the molecular and cellular responses of glia, or at least a subset thereof, will be directed to a different biological process than one expects in a clinical setting. Similarly, declines in adult neurogenesis have been associated with onset of dementia in AD [227]. This has led to hypotheses that decreased neurogenesis is causal for learning deficits and cognitive decline. Zebrafish represent an exciting opportunity to understand adult neurogenesis, however their continued robust generation of new neurons into adulthood directly opposes the disease process observed during clinical neurodegenerative decline. It will be of interest to examine if such neurogenesis is similarly affected in various zebrafish models of human neurodegeneration. We expect that it will be important to develop methods that allow adult neurogenesis to be slowed in these zebrafish disease models. A further complication to these disease models is that the diseaseassociated molecule being deployed in transgenic models is often intimately linked to proliferation. One example is amyloid precursor protein (APP), which is processed inappropriately in Alzheimer Disease and overexpressed in various Alzheimer Disease animal models. APP is enriched in neurogenic zones and acts to negatively modulate neurogenesis via its intracellular domain [228,229] and promotes neuritogenesis through its extracellular domains [230– 232]. Consistent with this, mouse models of Alzheimer Disease integrating APP, Tau and Presenilin have reduced adult neurogenesis [233–237]. Similarly, the prion protein (PrPC), often overexpressed in animal models to allow cross-species infectivity and disease modelling, is also involved in modulating neurogenesis. PrPC is expressed in adult neural precursors of mouse and positively modulates adult neurogenesis [238]. Overall, the substantial adult neurogenesis in zebrafish may be of great benefit to understanding the role for proteins like APP and PrP in neurogenesis in healthy brains. Degeneration models based on overexpression of disease-associated molecules may be especially prone to adult neural regeneration confounding data interpretations of described above. This will be especially true when the products that recapitulate disease phenotypes also inadvertently affect a proliferative response. We have briefly outlined potential confounds regarding interpretation and experimental design that are worth considering as zebrafish models of neurodegeneration become available. Nevertheless, in V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 each case, these confounds can be viewed as an opportunity to examine the differences across taxa that permit the adult brain to host proliferation leading to replacement of new neurons. For example, adult neurogenesis can be viewed as a therapeutic target for Alzheimer Disease treatments [239]. It is also noteworthy that stem cell therapy of degenerative diseases would need to be able to navigate these same obstacles as well as be able to define mechanisms to limit proliferation from becoming detrimental [146,240]. Thus it will be of great benefit to define mechanisms to limit growth signals from inducing death and to limit proliferation from becoming detrimental itself through hyperproliferation and cancer. Finally, in the same fashion as the retina, optic nerve and spinal cord efforts reviewed above, regeneration from intrinsic stem cells in the brain centres of zebrafish will offer opportunity to understand cellular and molecular events requisite to replacing neurons lost during disease progression. 5. Conclusion: zebrafish is a potent model of vertebrate CNS regeneration We have reviewed the intrinsic and impressive ability of zebrafish to regenerate CNS cells and tissues. Combined with a substantial repertoire of genetic tools and robust assays for functional integration of regenerated cells, zebrafish are uniquely positioned to allow design of sophisticated experiments to test hypotheses of CNS regeneration. These approaches can be complemented by high-throughput in vivo screens to explore for genetic and pharmacological modifiers of regeneration. Transgenic techniques and availability of these fish from stock centres are allowing refined cell ablation techniques that will herald cell-specific investigations of regeneration at both the molecular and functional levels. Finally, we note that several amenable paradigms are available to assess function of the regenerated CNS tissue, which ought be utilized to assess the integration of neurons into regenerated CNS circuits. Acknowledgments This work was funded by an NSERC Discovery Grant to WTA. VCF was supported by an SNF Fellowship and BF by an NSERC Scholarship. References [1] C.G. Gross, Neurogenesis in the adult brain: death of a dogma, Nat. Rev. Neurosci. 1 (1) (2000) 67–73. [2] D.C. Lie, et al., Neurogenesis in the adult brain: new strategies for central nervous system diseases, Annu. Rev. Pharmacol. Toxicol. 44 (2004) 399–421. [3] G. Kempermann, F.H. Gage, New nerve cells for the adult brain, Sci. Am. 280 (5) (1999) 48–53. [4] G.L. Ming, H. Song, Adult neurogenesis in the mammalian central nervous system, Annu. Rev. Neurosci. 28 (2005) 223–250. [5] A. Sanchez Alvarado, P.A. Tsonis, Bridging the regeneration gap: genetic insights from diverse animal models, Nat. Rev. Genet. 7 (11) (2006) 873–884. [6] S.L. Johnson, J.A. Weston, Temperature-sensitive mutations that cause stagespecific defects in zebrafish fin regeneration, Genetics 141 (4) (1995) 1583–1595. [7] R. Dahm, R. Geisler, Learning from small fry: the zebrafish as a genetic model organism for aquaculture fish species, Mar. Biotechnol. (N.Y.) 8 (4) (2006) 329–345. [8] S.C. Ekker, Morphants: a new systematic vertebrate functional genomics approach, Yeast 17 (4) (2000) 302–306. [9] Y. Doyon, et al., Heritable targeted gene disruption in zebrafish using designed zinc-finger nucleases, Nat. Biotechnol. 26 (6) (2008) 702–708. [10] X. Meng, et al., Targeted gene inactivation in zebrafish using engineered zincfinger nucleases, Nat. Biotechnol. 26 (6) (2008) 695–701. [11] Y.V. Makhankov, O. Rinner, S.C. Neuhauss, An inexpensive device for noninvasive electroretinography in small aquatic vertebrates, J. Neurosci. Meth. 135 (1–2) (2004) 205–210. [12] V.C. Fleisch, S.C. Neuhauss, Visual behavior in zebrafish, Zebrafish 3 (2) (2006) 191–201. [13] P.R. Johns, Growth of the adult goldfish eye. III. Source of the new retinal cells, J. Comp. Neurol. 176 (3) (1977) 343–357. [14] D.C. Otteson, P.F. Hitchcock, Stem cells in the teleost retina: persistent neurogenesis and injury-induced regeneration, Vision Res. 43 (8) (2003) 927–936. 377 [15] P.R. Johns, Formation of photoreceptors in larval and adult goldfish, J. Neurosci. 2 (2) (1982) 178–198. [16] P.A. Raymond, et al., Molecular characterization of retinal stem cells and their niches in adult zebrafish, BMC Dev. Biol. 6 (2006) 36. [17] D. Julian, K. Ennis, J.I. Korenbrot, Birth and fate of proliferative cells in the inner nuclear layer of the mature fish retina, J. Comp. Neurol. 394 (3) (1998) 271–282. [18] D.C. Otteson, A.R. D'Costa, P.F. Hitchcock, Putative stem cells and the lineage of rod photoreceptors in the mature retina of the goldfish, Dev. Biol. 232 (1) (2001) 62–76. [19] P. Hitchcock, et al., Persistent and injury-induced neurogenesis in the vertebrate retina, Prog. Retin. Eye Res. 23 (2) (2004) 183–194. [20] P.F. Hitchcock, et al., Antibodies against Pax6 immunostain amacrine and ganglion cells and neuronal progenitors, but not rod precursors, in the normal and regenerating retina of the goldfish, J. Neurobiol. 29 (3) (1996) 399–413. [21] E.M. Levine, et al., Restricted expression of a new paired-class homeobox gene in normal and regenerating adult goldfish retina, J. Comp. Neurol. 348 (4) (1994) 596–606. [22] S.A. Sullivan, et al., A goldfish Notch-3 homologue is expressed in neurogenic regions of embryonic, adult, and regenerating brain and retina, Dev. Genet. 20 (3) (1997) 208–223. [23] D.M. Wu, et al., Cones regenerate from retinal stem cells sequestered in the inner nuclear layer of adult goldfish retina, Invest. Ophthalmol. Vis. Sci. 42 (9) (2001) 2115–2124. [24] T.S. Vihtelic, D.R. Hyde, Light-induced rod and cone cell death and regeneration in the adult albino zebrafish (Danio rerio) retina, J. Neurobiol. 44 (3) (2000) 289–307. [25] T.S. Vihtelic, et al., Retinal regional differences in photoreceptor cell death and regeneration in light-lesioned albino zebrafish, Exp. Eye Res. 82 (4) (2006) 558–575. [26] D.D. Wang, A. Bordey, The astrocyte odyssey, Prog. Neurobiol. 86 (4) (2008) 342–367. [27] A. Bignami, et al., Localization of the glial fibrillary acidic protein in astrocytes by immunofluorescence, Brain Res. 43 (2) (1972) 429–435. [28] L.F. Eng, et al., An acidic protein isolated from fibrous astrocytes, Brain Res. 28 (2) (1971) 351–354. [29] J. Altman, Autoradiographic and histological studies of postnatal neurogenesis. IV. Cell proliferation and migration in the anterior forebrain, with special reference to persisting neurogenesis in the olfactory bulb, J. Comp. Neurol. 137 (4) (1969) 433–457. [30] F. Doetsch, et al., Subventricular zone astrocytes are neural stem cells in the adult mammalian brain, Cell 97 (6) (1999) 703–716. [31] M.B. Luskin, Restricted proliferation and migration of postnatally generated neurons derived from the forebrain subventricular zone, Neuron 11 (1) (1993) 173–189. [32] J. Altman, G.D. Das, Autoradiographic and histological evidence of postnatal hippocampal neurogenesis in rats, J. Comp. Neurol. 124 (3) (1965) 319–335. [33] H.A. Cameron, et al., Differentiation of newly born neurons and glia in the dentate gyrus of the adult rat, Neuroscience 56 (2) (1993) 337–344. [34] M.S. Kaplan, J.W. Hinds, Neurogenesis in the adult rat: electron microscopic analysis of light radioautographs, Science 197 (4308) (1977) 1092–1094. [35] A.J. Fischer, T.A. Reh, Muller glia are a potential source of neural regeneration in the postnatal chicken retina, Nat. Neurosci. 4 (3) (2001) 247–252. [36] A.J. Fischer, T.A. Reh, Potential of Muller glia to become neurogenic retinal progenitor cells, Glia 43 (1) (2003) 70–76. [37] R. Thummel, et al., Characterization of Muller glia and neuronal progenitors during adult zebrafish retinal regeneration, Exp. Eye Res. 87 (5) (2008) 433–444. [38] M.C. Senut, A. Gulati-Leekha, D. Goldman, An element in the alpha1-tubulin promoter is necessary for retinal expression during optic nerve regeneration but not after eye injury in the adult zebrafish, J. Neurosci. 24 (35) (2004) 7663–7673. [39] B.V. Fausett, D. Goldman, A role for alpha1 tubulin-expressing Muller glia in regeneration of the injured zebrafish retina, J. Neurosci. 26 (23) (2006) 6303–6313. [40] R.L. Bernardos, et al., Late-stage neuronal progenitors in the retina are radial Muller glia that function as retinal stem cells, J. Neurosci. 27 (26) (2007) 7028–7040. [41] A.C. Morris, T. Scholz, J.M. Fadool, Rod progenitor cells in the mature zebrafish retina, Adv. Exp. Med. Biol. 613 (2008) 361–368. [42] R. Thummel, et al., Inhibition of Muller glial cell division blocks regeneration of the light-damaged zebrafish retina, Dev. Neurobiol. 68 (3) (2008) 392–408. [43] S.M. Fimbel, et al., Regeneration of inner retinal neurons after intravitreal injection of ouabain in zebrafish, J. Neurosci. 27 (7) (2007) 1712–1724. [44] M.L. Chang, et al., Reactive changes of retinal astrocytes and Muller glial cells in kainate-induced neuroexcitotoxicity, J. Anat. 210 (1) (2007) 54–65. [45] M.A. Dyer, C.L. Cepko, Control of Muller glial cell proliferation and activation following retinal injury, Nat. Neurosci. 3 (9) (2000) 873–880. [46] X. Zhao, et al., Growth factor-responsive progenitors in the postnatal mammalian retina, Dev. Dyn. 232 (2) (2005) 349–358. [47] J.L. Close, et al., Epidermal growth factor receptor expression regulates proliferation in the postnatal rat retina, Glia 54 (2) (2006) 94–104. [48] A.V. Das, et al., Neural stem cell properties of Muller glia in the mammalian retina: regulation by Notch and Wnt signaling, Dev. Biol. 299 (1) (2006) 283–302. [49] F. Osakada, et al., Wnt signaling promotes regeneration in the retina of adult mammals, J. Neurosci. 27 (15) (2007) 4210–4219. [50] J. Wan, et al., Preferential regeneration of photoreceptor from Muller glia after retinal degeneration in adult rat, Vision Res. 48 (2) (2008) 223–234. 378 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 [51] J. Wan, et al., Sonic hedgehog promotes stem-cell potential of Muller glia in the mammalian retina, Biochem. Biophys. Res. Commun. 363 (2) (2007) 347–354. [52] B. Bhatia, et al., Distribution of Muller stem cells within the neural retina: evidence for the existence of a ciliary margin-like zone in the adult human eye, Exp. Eye Res. 89 (3) (2009) 373–382. [53] D.A. Cameron, et al., Gene expression profiles of intact and regenerating zebrafish retina, Mol. Vis. 11 (2005) 775–791. [54] A.A. Calinescu, et al., Cellular expression of midkine-a and midkine-b during retinal development and photoreceptor regeneration in zebrafish, J. Comp. Neurol. 514 (1) (2009) 1–10. [55] S. Craig, A.-A. Calinescu, P.F. Hitchcock, Identification of the molecular signatures integral to regenerating photoreceptors in the retina of the zebrafish, J. Ocul. Biol. 1 (2008) 73–84. [56] S.C. Kassen, et al., Time course analysis of gene expression during light-induced photoreceptor cell death and regeneration in albino zebrafish, Dev. Neurobiol. 67 (8) (2007) 1009–1031. [57] Z. Qin, L.K. Barthel, P.A. Raymond, Genetic evidence for shared mechanisms of epimorphic regeneration in zebrafish, Proc. Natl Acad. Sci. USA 106 (23) (2009) 9310–9315. [58] C.L. Lien, et al., Gene expression analysis of zebrafish heart regeneration, PLoS Biol. 4 (8) (2006) e260. [59] M. Schebesta, et al., Transcriptional profiling of caudal fin regeneration in zebrafish, Sci. World J. 6 (Suppl 1) (2006) 38–54. [60] S.C. Kassen, et al., CNTF induces photoreceptor neuroprotection and Muller glial cell proliferation through two different signaling pathways in the adult zebrafish retina, Exp. Eye Res. 88 (6) (2009) 1051–1064. [61] P. Yurco, D.A. Cameron, Cellular correlates of proneural and Notch–delta gene expression in the regenerating zebrafish retina, Vis. Neurosci. 24 (3) (2007) 437–443. [62] B.V. Fausett, J.D. Gumerson, D. Goldman, The proneural basic helix–loop–helix gene ascl1a is required for retina regeneration, J. Neurosci. 28 (5) (2008) 1109–1117. [63] A. Rattner, J. Nathans, The genomic response to retinal disease and injury: evidence for endothelin signaling from photoreceptors to glia, J. Neurosci. 25 (18) (2005) 4540–4549. [64] U.K. Hanisch, H. Kettenmann, Microglia: active sensor and versatile effector cells in the normal and pathologic brain, Nat. Neurosci. 10 (11) (2007) 1387–1394. [65] L. Benowitz, Y. Yin, Rewiring the injured CNS: lessons from the optic nerve, Exp. Neurol. 209 (2) (2008) 389–398. [66] B. Lorber, M. Berry, A. Logan, Lens injury stimulates adult mouse retinal ganglion cell axon regeneration via both macrophage- and lens-derived factors, Eur. J. Neurosci. 21 (7) (2005) 2029–2034. [67] T. Okada, et al., Intravitreal macrophage activation enables cat retinal ganglion cells to regenerate injured axons into the mature optic nerve, Exp. Neurol. 196 (1) (2005) 153–163. [68] V. Pernet, A. Di Polo, Synergistic action of brain-derived neurotrophic factor and lens injury promotes retinal ganglion cell survival, but leads to optic nerve dystrophy in vivo, Brain 129 (Pt 4) (2006) 1014–1026. [69] Y. Yin, et al., Macrophage-derived factors stimulate optic nerve regeneration, J. Neurosci. 23 (6) (2003) 2284–2293. [70] D. Fischer, Z. He, L.I. Benowitz, Counteracting the Nogo receptor enhances optic nerve regeneration if retinal ganglion cells are in an active growth state, J. Neurosci. 24 (7) (2004) 1646–1651. [71] D. Fischer, et al., Switching mature retinal ganglion cells to a robust growth state in vivo: gene expression and synergy with RhoA inactivation, J. Neurosci. 24 (40) (2004) 8726–8740. [72] R.R. Bernhardt, Cellular and molecular bases of axonal regeneration in the fish central nervous system, Exp. Neurol. 157 (2) (1999) 223–240. [73] C.G. Becker, T. Becker, Growth and pathfinding of regenerating axons in the optic projection of adult fish, J. Neurosci. Res. 85 (12) (2007) 2793–2799. [74] C.G. Becker, T. Becker, Adult zebrafish as a model for successful central nervous system regeneration, Restor. Neurol. Neurosci. 26 (2–3) (2008) 71–80. [75] L.I. Benowitz, A. Routtenberg, GAP-43: an intrinsic determinant of neuronal development and plasticity, Trends Neurosci. 20 (2) (1997) 84–91. [76] M. Kaneda, et al., Changes of phospho-growth-associated protein 43 (phosphoGAP43) in the zebrafish retina after optic nerve injury: a long-term observation, Neurosci. Res. 61 (3) (2008) 281–288. [77] J.A. Miotke, A.J. MacLennan, R.L. Meyer, Immunohistochemical localization of CNTFRalpha in adult mouse retina and optic nerve following intraorbital nerve crush: evidence for the axonal loss of a trophic factor receptor after injury, J. Comp. Neurol. 500 (2) (2007) 384–400. [78] A. Muller, T.G. Hauk, D. Fischer, Astrocyte-derived CNTF switches mature RGCs to a regenerative state following inflammatory stimulation, Brain 130 (Pt 12) (2007) 3308–3320. [79] A. Muller, et al., Exogenous CNTF stimulates axon regeneration of retinal ganglion cells partially via endogenous CNTF, Mol. Cell. Neurosci. 41 (2) (2009) 233–246. [80] B. Petrausch, et al., A purine-sensitive pathway regulates multiple genes involved in axon regeneration in goldfish retinal ganglion cells, J. Neurosci. 20 (21) (2000) 8031–8041. [81] C. Munderloh, et al., Reggies/flotillins regulate retinal axon regeneration in the zebrafish optic nerve and differentiation of hippocampal and N2a neurons, J. Neurosci. 29 (20) (2009) 6607–6615. [82] C.A. Stuermer, The reggie/flotillin connection to growth, Trends Cell Biol. 20 (1) (2009) 6–13. [83] M.B. Veldman, et al., Gene expression analysis of zebrafish retinal ganglion cells during optic nerve regeneration identifies KLF6a and KLF7a as important regulators of axon regeneration, Dev. Biol. 312 (2) (2007) 596–612. [84] D.L. Moore, et al., KLF family members regulate intrinsic axon regeneration ability, Science 326 (5950) (2009) 298–301. [85] J. Schweitzer, et al., Expression of protein zero is increased in lesioned axon pathways in the central nervous system of adult zebrafish, Glia 41 (3) (2003) 301–317. [86] J. Schweitzer, et al., Contactin1a expression is associated with oligodendrocyte differentiation and axonal regeneration in the central nervous system of zebrafish, Mol. Cell. Neurosci. 35 (2) (2007) 194–207. [87] T. Matsukawa, et al., Role of purpurin as a retinol-binding protein in goldfish retina during the early stage of optic nerve regeneration: its priming action on neurite outgrowth, J. Neurosci. 24 (38) (2004) 8346–8353. [88] M. Nagashima, et al., Involvement of retinoic acid signaling in goldfish optic nerve regeneration, Neurochem. Int. 54 (3–4) (2009) 229–236. [89] B. Petrausch, et al., Lesion-induced regulation of netrin receptors and modification of netrin-1 expression in the retina of fish and grafted rats, Mol. Cell. Neurosci. 16 (4) (2000) 350–364. [90] J.W. Fawcett, Overcoming inhibition in the damaged spinal cord, J. Neurotrauma 23 (3–4) (2006) 371–383. [91] T. Sivron, M. Schwartz, Nonpermissive nature of fish optic nerves to axonal growth is due to presence of myelin-associated growth inhibitors, Exp. Neurol. 130 (2) (1994) 411–413. [92] M. Wanner, et al., Reevaluation of the growth-permissive substrate properties of goldfish optic nerve myelin and myelin proteins, J. Neurosci. 15 (11) (1995) 7500–7508. [93] M. Bastmeyer, et al., Growth of regenerating goldfish axons is inhibited by rat oligodendrocytes and CNS myelin but not but not by goldfish optic nerve tract oligodendrocytelike cells and fish CNS myelin, J. Neurosci. 11 (3) (1991) 626–640. [94] M.E. Schwab, Nogo and axon regeneration, Curr. Opin. Neurobiol. 14 (1) (2004) 118–124. [95] Y. Su, et al., Axonal regeneration after optic nerve crush in Nogo-A/B/C knockout mice, Mol. Vis. 14 (2008) 268–273. [96] H. Diekmann, et al., Analysis of the reticulon gene family demonstrates the absence of the neurite growth inhibitor Nogo-A in fish, Mol. Biol. Evol. 22 (8) (2005) 1635–1648. [97] H. Abdesselem, et al., No Nogo66- and NgR-mediated inhibition of regenerating axons in the zebrafish optic nerve, J. Neurosci. 29 (49) (2009) 15489–15498. [98] J. Silver, J.H. Miller, Regeneration beyond the glial scar, Nat. Rev. Neurosci. 5 (2) (2004) 146–156. [99] P. Charalambous, L.A. Hurst, S. Thanos, Engrafted chicken neural tube-derived stem cells support the innate propensity for axonal regeneration within the rat optic nerve, Invest. Ophthalmol. Vis. Sci. 49 (8) (2008) 3513–3524. [100] C.G. Becker, T. Becker, Repellent guidance of regenerating optic axons by chondroitin sulfate glycosaminoglycans in zebrafish, J. Neurosci. 22 (3) (2002) 842–853. [101] A. Shirvan, et al., Anti-semaphorin 3A antibodies rescue retinal ganglion cells from cell death following optic nerve axotomy, J. Biol. Chem. 277 (51) (2002) 49799–49807. [102] D.C. Callander, et al., Expression of multiple class three semaphorins in the retina and along the path of zebrafish retinal axons, Dev. Dyn. 236 (10) (2007) 2918–2924. [103] C.G. Becker, et al., Tenascin-R as a repellent guidance molecule for developing optic axons in zebrafish, J. Neurosci. 23 (15) (2003) 6232–6237. [104] C.G. Becker, et al., Tenascin-R as a repellent guidance molecule for newly growing and regenerating optic axons in adult zebrafish, Mol. Cell. Neurosci. 26 (3) (2004) 376–389. [105] T. Becker, et al., Tenascin-R inhibits regrowth of optic fibers in vitro and persists in the optic nerve of mice after injury, Glia 29 (4) (2000) 330–346. [106] C.G. Becker, R.L. Meyer, T. Becker, Gradients of ephrin-A2 and ephrin-A5b mRNA during retinotopic regeneration of the optic projection in adult zebrafish, J. Comp. Neurol. 427 (3) (2000) 469–483. [107] J. Rodger, et al., EphA/ephrin-A interactions during optic nerve regeneration: restoration of topography and regulation of ephrin-A2 expression, Mol. Cell. Neurosci. 25 (1) (2004) 56–68. [108] C.E. King, et al., Transient up-regulation of retinal EphA3 and EphA5, but not ephrin-A2, coincides with re-establishment of a topographic map during optic nerve regeneration in goldfish, Exp. Neurol. 183 (2) (2003) 593–599. [109] M. Nagashima, et al., Purpurin is a key molecule for cell differentiation during the early development of zebrafish retina, Brain Res. (2009). [110] P. Pesheva, et al., The F3/11 cell adhesion molecule mediates the repulsion of neurons by the extracellular matrix glycoprotein J1-160/180, Neuron 10 (1) (1993) 69–82. [111] K.S. Cho, D.F. Chen, Promoting optic nerve regeneration in adult mice with pharmaceutical approach, Neurochem. Res. 33 (10) (2008) 2126–2133. [112] K.S. Cho, et al., Re-establishing the regenerative potential of central nervous system axons in postnatal mice, J. Cell Sci. 118 (Pt 5) (2005) 863–872. [113] A.R. Harvey, M. Hellstrom, J. Rodger, Gene therapy and transplantation in the retinofugal pathway, Prog. Brain Res. 175 (2009) 151–161. [114] M.P. Callahan, A.F. Mensinger, Restoration of visual function following optic nerve regeneration in bluegill (Lepomis macrochirus) × pumpkinseed (Lepomis gibbosus) hybrid sunfish, Vis. Neurosci. 24 (3) (2007) 309–317. [115] R.A. Meyer, M.E. Zaruba, G.M. McKhann, Flow-cytometry of isolated cells from the brain, Anal. Quant. Cytol. Histol. 2 (1) (1980) 66–74. [116] R.L. Meyer, Eye-in-water electrophysiological mapping of goldfish with and without tectal lesions, Exp. Neurol. 56 (1) (1977) 23–41. [117] D.P. Northmore, T. Masino, Recovery of vision in fish after optic nerve crush: a behavioral and electrophysiological study, Exp. Neurol. 84 (1) (1984) 109–125. V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 [118] C.G. Becker, T. Becker, Model Organisms in Spinal Cord Regeneration, Wiley-VCH Verlag GmbH, 2007. [119] P. Chapouton, R. Jagasia, L. Bally-Cuif, Adult neurogenesis in non-mammalian vertebrates, Bioessays 29 (8) (2007) 745–757. [120] A. Blesch, M.H. Tuszynski, Spinal cord injury: plasticity, regeneration and the challenge of translational drug development, Trends Neurosci. 32 (1) (2009) 41–47. [121] A.F. Cristante, et al., Stem cells in the treatment of chronic spinal cord injury: evaluation of somatosensitive evoked potentials in 39 patients, Spinal Cord 47 (10) (2009) 733–738. [122] G.W.J. Hawryluk, et al., Protection and repair of the injured spinal cord: a review of completed, ongoing, and planned clinical trials for acute spinal cord injury, Neurosurg. Focus 25 (5) (2008). [123] J.W. Rowland, et al., Current status of acute spinal cord injury pathophysiology and emerging therapies: promise on the horizon, Neurosurg. Focus 25 (5) (2008). [124] T. Becker, et al., Axonal regrowth after spinal cord transection in adult zebrafish, J. Comp. Neurol. 377 (4) (1997) 577–595. [125] T. Becker, et al., Readiness of zebrafish brain neurons to regenerate a spinal axon correlates with differential expression of specific cell recognition molecules, J. Neurosci. 18 (15) (1998) 5789–5803. [126] T. Becker, et al., Differences in the regenerative response of neuronal cell populations and indications for plasticity in intraspinal neurons after spinal cord transection in adult zebrafish, Mol. Cell. Neurosci. 30 (2) (2005) 265–278. [127] C.G. Becker, et al., L1.1 is involved in spinal cord regeneration in adult zebrafish, J. Neurosci. 24 (36) (2004) 7837–7842. [128] J. Haspel, M. Grumet, The L1CAM extracellular region: a multi-domain protein with modular and cooperative binding modes, Front. Biosci. 8 (2003) S1210–S1225. [129] C. Roonprapunt, et al., Soluble cell adhesion molecule L1-fc promotes locomotor recovery in rats after spinal cord injury, J. Neurotrauma 20 (9) (2003) 871–882. [130] C.M. Booth, M.C. Brown, Expression of GAP-43 mRNA in mouse spinal cord following unilateral peripheral nerve damage: is there a contralateral effect? Eur. J. Neurosci. 5 (12) (1993) 1663–1676. [131] H. Linda, et al., Expression of GAP-43 mRNA in the adult mammalian spinal cord under normal conditions and after different types of lesions, with special reference to motoneurons, Exp. Brain Res. 91 (2) (1992) 284–295. [132] A.B. Schmitt, et al., GAP-43 (B-50) and C-Jun are up-regulated in axotomized neurons of Clarke's nucleus after spinal cord injury in the adult rat, Neurobiol. Dis. 6 (2) (1999) 122–130. [133] G. Jeserich, K. Klempahn, M. Pfeiffer, Features and functions of oligodendrocytes and myelin proteins of lower vertebrate species, J. Mol. Neurosci. 35 (1) (2008) 117–126. [134] P. Drapeau, et al., In vivo recording from identifiable neurons of the locomotor network in the developing zebrafish, J. Neurosci. Meth. 88 (1) (1999) 1–13. [135] J.I.P. Prugh, C.B. Kimmel, W.K. Metcalfe, Non-invasive recording of the Mauthner neuron action potential in larval zebrafish, J. Exp. Biol. 101 (DEC 1982) 83–92. [136] M. Chong, P. Drapeau, Interaction between hindbrain and spinal networks during the development of locomotion in zebrafish, Dev. Neurobiol. 67 (7) (2007) 933–947. [137] J.R. Fetcho, The utility of zebrafish for studies of the comparative biology of motor systems, J. Exp. Zool. B Mol. Dev. Evol. 308B (5) (2007) 550–562. [138] D.H. Bhatt, et al., Cyclic AMP-induced repair of zebrafish spinal circuits, Science 305 (5681) (2004) 254–258. [139] D. Kokel, R.T. Peterson, Chemobehavioural phenomics and behaviour-based psychiatric drug discovery in the zebrafish, Brief. Funct. Genomic. Proteomic. 7 (6) (2008) 483–490. [140] L.K. Mathew, et al., Unraveling tissue regeneration pathways using chemical genetics, J. Biol. Chem. 282 (48) (2007) 35202–35210. [141] R.T. Peterson, et al., Use of non-mammalian alternative models for neurotoxicological study, Neurotoxicology 29 (3) (2008) 546–555. [142] T. Becker, C.G. Becker, Regenerating descending axons preferentially reroute to the gray matter in the presence of a general macrophage/microglial reaction caudal to a spinal transection in adult zebrafish, J. Comp. Neurol. 433 (1) (2001) 131–147. [143] J. Frisen, et al., Adhesive/repulsive properties in the injured spinal cord: relation to myelin phagocytosis by invading macrophages, Exp. Neurol. 129 (2) (1994) 183–193. [144] S.U. Kim, J. de Vellis, Stem cell-based cell therapy in neurological diseases: a review, J. Neurosci. Res. 87 (10) (2009) 2183–2200. [145] K. Sugaya, et al., Practical issues in stem cell therapy for Alzheimer's disease, Curr. Alzheimer Res. 4 (4) (2007) 370–377. [146] P. Taupin, Adult neurogenesis, neural stem cells and Alzheimer's Disease: developments, limitations, problems and promises, Curr. Alzheimer Res. 6 (6) (2009) 461–470. [147] M. Valenzuela, et al., Neural stem cell therapy for neuropsychiatric disorders, Acta Neuropsychiatr. 19 (1) (2007) 11–26. [148] H. Grandel, et al., Neural stem cells and neurogenesis in the adult zebrafish brain: origin, proliferation dynamics, migration and cell fate, Dev. Biol. 295 (1) (2006) 263–277. [149] J. Kaslin, et al., Stem cells in the adult zebrafish cerebellum: initiation and maintenance of a novel stem cell niche, J. Neurosci. 29 (19) (2009) 6142–6153. [150] M.F. Wulliman, Neuroanatomy of Zebrafish Brain, Birkhauser Boston, 1996. [151] G.K.H. Zupanc, Neurogenesis and neuronal regeneration in the adult fish brain, J. Comp. Physiol. Neuroethol. Sens. Neural Behav. Physiol. 192 (6) (2006) 649–670. [152] G.K.H. Zupanc, Towards brain repair: insights from teleost fish, Semin. Cell Dev. Biol. 20 (6) (2009) 683–690. 379 [153] G.K.H. Zupanc, K. Hinsch, F.H. Gage, Proliferation, migration, neuronal differentiation, and long-term survival of new cells in the adult zebrafish brain, J. Comp. Neurol. 488 (3) (2005) 290–319. [154] K. Hinsch, G.K.H. Zupanc, Isolation, cultivation, and differentiation of neural stem cells from adult fish brain, J. Neurosci. Meth. 158 (1) (2006) 75–88. [155] K. Hinsch, G.K.H. Zupanc, Generation and long-term persistence of new neurons in the adult zebrafish brain: a quantitative analysis, Neuroscience 146 (2) (2007) 679–696. [156] B. Adolf, et al., Conserved and acquired features of adult neurogenesis in the zebrafish telencephalon, Dev. Biol. 295 (1) (2006) 278–293. [157] V. Kroehne, J. Kaslin, M. Brand, Proliferation and cell fates during regeneration of the adult zebrafish brain, Mech. Dev. 126 (2009) S291–S292. [158] G.K. Zupanc, et al., Generation, long-term persistence, and neuronal differentiation of cells with nuclear aberrations in the adult zebrafish brain, Neuroscience 159 (4) (2009) 1338–1348. [159] P. Chapouton, et al., her5 expression reveals a pool of neural stem cells in the adult zebrafish midbrain, Development 133 (21) (2006) 4293–4303. [160] C. Stigloher, et al., Identification of Neural Progenitor Pools by E(Spl) Factors in the Embryonic and Adult Brain, 2008. [161] J.E. Braisted, T.F. Essman, P.A. Raymond, Selective regeneration of photoreceptors in goldfish retina, Development 120 (9) (1994) 2409–2419. [162] S.G. He, R.H. Masland, Retinal direction selectivity after targeted laser ablation of starburst amacrine cells, Nature 389 (6649) (1997) 378–382. [163] K.S. Liu, J.R. Fetcho, Laser ablations reveal functional relationships of segmental hindbrain neurons in zebrafish, Neuron 23 (2) (1999) 325–335. [164] W. Maier, H. Wolburg, Regeneration of the goldfish retina after exposure to different doses of ouabain, Cell Tissue Res. 202 (1) (1979) 99–118. [165] S. Curado, D.Y.R. Stainier, R.M. Anderson, Nitroreductase-mediated cell/tissue ablation in zebrafish: a spatially and temporally controlled ablation method with applications in developmental and regeneration studies, Nat. Protoc. 3 (6) (2008) 948–954. [166] N.C. Gellrich, et al., Quantification of histological changes after calibrated crush of the intraorbital optic nerve in rats, Br. J. Ophthalmol. 86 (2) (2002) 233–237. [167] L. Sarikcioglu, N. Demir, A. Demirtop, A standardized method to create optic nerve crush: Yasargil aneurysm clip, Exp. Eye Res. 84 (2) (2007) 373–377. [168] G. Alonso, A. Privat, Neuropeptide Y-producing neurons of the arcuate nucleus regenerate axons after surgical deafferentation of the mediobasal hypothalamus, J. Neurosci. Res. 34 (5) (1993) 510–522. [169] E.J. Bradbury, et al., Chondroitinase ABC promotes functional recovery after spinal cord injury, Nature 416 (6881) (2002) 636–640. [170] S.J. Zottoli, M.M. Freemer, Recovery of C-starts, equilibrium and targeted feeding after whole spinal cord crush in the adult goldfish Carassius auratus, J. Exp. Biol. 206 (17) (2003) 3015–3029. [171] M. Schwartz, Optic nerve crush: protection and regeneration, Brain Res. Bull. 62 (6) (2004) 467–471. [172] M. Bilgen, Magnetic resonance microscopy of spinal cord injury in mouse using a miniaturized implantable RF coil, J. Neurosci. Meth. 159 (1) (2007) 93–97. [173] J.M.W. Slack, G. Lin, Y. Chen, The Xenopus tadpole: a new model for regeneration research, Cell. Mol. Life Sci. 65 (1) (2008) 54–63. [174] G.K. Zupanc, Adult neurogenesis and neuronal regeneration in the brain of teleost fish, J. Physiol. Paris 102 (4–6) (2008) 357–373. [175] R.L. Meyer, Mapping the normal and regenerating retinotectal projection of goldfish with autoradiographic methods, J. Comp. Neurol. 189 (2) (1980) 273–289. [176] D.A. Cameron, S.S. Easter, Cone photoreceptor regeneration in adult fish retina — phenotypic determination and mosaic pattern formation, J. Neurosci. 15 (3) (1995) 2255–2271. [177] P.F. Hitchcock, et al., Local regeneration in the retina of the goldfish, J. Neurobiol. 23 (2) (1992) 187–203. [178] A.F. Mensinger, M.K. Powers, Visual function in regenerating teleost retina following surgical lesioning, Vis. Neurosci. 24 (3) (2007) 299–307. [179] G. Fridberg, et al., Regeneration of caudal neurosecretory system in cichlid teleost Tilapia mossambica, J. Exp. Zool. 162 (3) (1966) 311–366. [180] D.S. Hughes, R.R. Schade, M. Ontell, Ablation of the fetal mouse spinal cord — the effect on soleus muscle, Dev. Dyn. 193 (2) (1992) 164–174. [181] T. Roeser, H. Baier, Visuomotor behaviors in larval zebrafish after GFP-guided laser ablation of the optic tectum, J. Neurosci. 23 (9) (2003) 3726–3734. [182] C.B. Kimmel, K. Hatta, W.K. Metcalfe, Early axonal contacts during development of an identified dendrite in the brain of the zebrafish, Neuron 4 (4) (1990) 535–545. [183] V. Kohli, A.Y. Elezzabi, Laser surgery of zebrafish (Danio rerio) embryos using femtosecond laser pulses: optimal parameters for exogenous material delivery, and the laser's effect on short- and long-term development, BMC Biotechnol. 8 (2008) 7. [184] M.F. Yanik, et al., Neurosurgery: functional regeneration after laser axotomy, Nature 432 (7019) (2004) 822. [185] S. Shahinfar, D.P. Edward, M.O.M. Tso, A pathological study of photoreceptor cell death in retinal photic injury, Curr. Eye Res. 10 (1) (1991) 47–59. [186] L.R. Marotte, J. Wyedvorak, R.F. Mark, Retinotectal reorganization in goldfish. 2. Effects of partial tectal ablation and constant light on the retina, Neuroscience 4 (6) (1979) 803–810. [187] D.T. Organisciak, B.S. Winkler, Retinal light damage — practical and theoretical considerations, Prog. Retin. Eye Res. 13 (1) (1994) 1–29. [188] J.S. Penn, Effects of continous light on the retina of a fish, notemigonuscrysoleucas, J. Comp. Neurol. 238 (1) (1985) 121–127. 380 V.C. Fleisch et al. / Biochimica et Biophysica Acta 1812 (2011) 364–380 [189] A. Wenzel, et al., Molecular mechanisms of light-induced photoreceptor apoptosis and neuroprotection for retinal degeneration, Prog. Retin. Eye Res. 24 (2) (2005) 275–306. [190] A.S. Abler, et al., Photic injury triggers apoptosis of photoreceptor cells, Res. Commun. Mol. Pathol. Pharmacol. 92 (2) (1996) 177–189. [191] W.T. Allison, et al., Degeneration and regeneration of ultraviolet cone photoreceptors during development in rainbow trout, J. Comp. Neurol. 499 (5) (2006) 702–715. [192] P.A. Raymond, M.J. Reifler, P.K. Rivlin, Regeneration of goldfish retina: rod precursors are a likely source of regenerated cells, J. Neurobiol. 19 (5) (1988) 431–463. [193] R.C. Valente, et al., Mechanisms of ouabain toxicity, FASEB J. 17 (10) (2003) 1700–1702. [194] T. Sherpa, et al., Ganglion cell regeneration following whole-retina destruction in zebrafish, Dev. Neurobiol. 68 (2) (2008) 166–181. [195] L. Li, J.E. Dowling, Effects of dopamine depletion on visual sensitivity of zebrafish, J. Neurosci. 20 (5) (2000) 1893–1903. [196] Z.S. Lin, S. Yazulla, Depletion of retinal dopamine does not affect the ERG b-wave increment threshold function in goldfish in vivo, Vis. Neurosci. 11 (4) (1994) 695–702. [197] Z.S. Lin, S. Yazulla, Depletion of retinal dopamine increases brightness perception in goldfish, Vis. Neurosci. 11 (4) (1994) 683–693. [198] K. Negishi, T. Teranishi, S. Kato, New dopaminergic and indoleamine-accumulating cells in the growth zone of goldfish retinas after neurotoxic destruction, Science 216 (4547) (1982) 747–749. [199] K. Negishi, T. Teranishi, S. Kato, Growth zone of the juvenile goldfish retina revealed by fluorescent flat mounts, J. Neurosci. Res. 7 (3) (1982) 321–330. [200] A.C. Morris, et al., Cone survival despite rod degeneration in XOPS-mCFP transgenic zebrafish, Invest. Ophthalmol. Vis. Sci. 46 (12) (2005) 4762–4771. [201] G. Stearns, et al., A mutation in the cone-specific pde6 gene causes rapid cone photoreceptor degeneration in zebrafish, J. Neurosci. 27 (50) (2007) 13866–13874. [202] A.C. Morris, et al., Genetic dissection reveals two separate pathways for rod and cone regeneration in the teleost retina, Dev. Neurobiol. 68 (5) (2008) 605–619. [203] J.M. Davison, et al., Transactivation from Gal4-VP16 transgenic insertions for tissuespecific cell labeling and ablation in zebrafish, Dev. Biol. 304 (2) (2007) 811–824. [204] X.F. Zhao, S. Ellingsen, A. Fjose, Labelling and targeted ablation of specific bipolar cell types in the zebrafish retina, BMC Neurosci. 10 (2009). [205] H. Pisharath, Validation of nitroreductase, a prodrug-activating enzyme, mediated cell death in embryonic zebrafish (Danio rerio), Comp. Med. 57 (3) (2007) 241–246. [206] K. Asakawa, et al., Genetic dissection of neural circuits by Tol2 transposonmediated Gal4 gene and enhancer trapping in zebrafish, Proc. Natl Acad. Sci. USA 105 (4) (2008) 1255–1260. [207] E.K. Scott, H. Baier, The cellular architecture of the larval zebrafish tectum, as revealed by gal4 enhancer trap lines, Front. Neural Circuits 3 (2009) 13. [208] E.K. Scott, et al., Targeting neural circuitry in zebrafish using GAL4 enhancer trapping, Nat. Methods 4 (4) (2007) 323–326. [209] S. Saszik, J. Bilotta, C.M. Givin, ERG assessment of zebrafish retinal development, Vis. Neurosci. 16 (5) (1999) 881–888. [210] M.W. Seeliger, et al., Genetic separation of rod and cone-driven activity in RPE65 knockout mice identifies the rods as the source of vision in a form of lebers congenital amaurosis, Invest. Ophthalmol. Vis. Sci. 43 (2002) 3682. [211] J.E. Dowling, Functional and pharmacological organization of the retina, Arch. Neurol. 44 (11) (1987) 1208-1208. [212] A.F. Mensinger, M.K. Powers, Visual function in regenerating teleost retina following cytotoxic lesioning, Vis. Neurosci. 16 (2) (1999) 241–251. [213] H.I. Browman, C.W. Hawryshyn, The developmental trajectory of ultraviolet photosensitivity in rainbow trout is altered by thyroxine, Vision Res. 34 (11) (1994) 1397–1406. [214] J.P. Gabriel, et al., Locomotor pattern in the adult zebrafish spinal cord in vitro, J. Neurophysiol. 99 (1) (2008) 37–48. [215] W. van Raamsdonk, et al., Long term effects of spinal cord transection in zebrafish: swimming performances, and metabolic properties of the neuromuscular system, Acta Histochem. 100 (2) (1998) 117–131. [216] S.C.F. Neuhauss, Behavioral genetic approaches to visual system development and function in zebrafish, J. Neurobiol. 54 (1) (2003) 148–160. [217] S.E. Brockerhoff, Measuring the optokinetic response of zebrafish larvae, Nat. Protoc. 1 (5) (2006) 2448–2451. [218] S.S. Easter, G.N. Nicola, The development of eye movements in the zebrafish (Danio rerio), Dev. Psychobiol. 31 (4) (1997) 267–276. [219] O. Rinner, J.M. Rick, S.C.F. Neuhauss, Contrast sensitivity, spatial and temporal tuning of the larval zebrafish optokinetic response, Invest. Ophthalmol. Vis. Sci. 46 (1) (2005) 137–142. [220] Y.Y. Huang, S.C. Neuhauss, The optokinetic response in zebrafish and its applications, Front. Biosci. 13 (2008) 1899–1916. [221] A.E. Lindsey, M.K. Powers, Visual behavior of adult goldfish with regenerating retina, Vis. Neurosci. 24 (3) (2007) 247–255. [222] S.C.F. Neuhauss, et al., Genetic disorders of vision revealed by a behavioral screen of 400 essential loci in zebrafish, J. Neurosci. 19 (19) (1999) 8603–8615. [223] M.B. Orger, H. Baier, Channeling of red and green cone inputs to the zebrafish optomotor response, Vis. Neurosci. 22 (3) (2005) 275–281. [224] A.L. Phinney, et al., Mouse models of Alzheimer's disease: the long and filamentous road, Neurol. Res. 25 (6) (2003) 590–600. [225] W.T. Allison, et al., Photic history modifies susceptibility to retinal damage in albino trout, Vis. Neurosci. 23 (1) (2006) 25–34. [226] E. Salta, et al., Evaluation of the possible transmission of BSE and scrapie to gilthead sea bream (Sparus aurata), PLoS ONE 4 (7) (2009) e6175. [227] R.J.E. Armstrong, R.A. Barker, Neurodegeneration: a failure of neuroregeneration? Lancet 358 (9288) (2001) 1174–1176. [228] Q.H. Ma, et al., A TAG1-APP signalling pathway through Fe65 negatively modulates neurogenesis, Nat. Cell Biol. 10 (3) (2008) 283–294. [229] P. Porayette, et al., Differential processing of amyloid-beta precursor protein directs human embryonic stem cell proliferation and differentiation into neuronal precursor cells, J. Biol. Chem. 284 (35) (2009) 23806–23817. [230] Y. Chen, C. Dong, A beta 40 promotes neuronal cell fate in neural progenitor cells, Cell Death Differ. 16 (3) (2009) 386–394. [231] M. Gralle, S.T. Ferreira, Structure and functions of the human amyloid precursor protein: the whole is more than the sum of its parts, Prog. Neurobiol. 82 (1) (2007) 11–32. [232] E.A. Milward, et al., The amyloid protein precursor of Alzheimer disease is a mediator of the effects of nerve growth factor on neurite outgrowth, Neuron 9 (1) (1992) 129–137. [233] S.H. Choi, et al., Non-cell-autonomous effects of presenilin 1 variants on enrichment-mediated hippocampal progenitor cell proliferation and differentiation, Neuron 59 (4) (2008) 568–580. [234] F.V. Ermini, et al., Neurogenesis and alterations of neural stem cells in mouse models of cerebral amyloidosis, Am. J. Pathol. 172 (6) (2008) 1520–1528. [235] J.J. Rodriguez, V.C. Jones, A. Verkhratsky, Impaired cell proliferation in the subventricular zone in an Alzheimer's disease model, NeuroReport 20 (10) (2009) 907–912. [236] P.H. Wen, et al., Overexpression of wild type but not an FAD mutant presenilin-1 promotes neurogenesis in the hippocampus of adult mice, Neurobiol. Dis. 10 (1) (2002) 8–19. [237] C. Zhang, et al., Long-lasting impairment in hippocampal neurogenesis associated with amyloid deposition in a knock-in mouse model of familial Alzheimer's disease, Exp. Neurol. 204 (1) (2007) 77–87. [238] A.D. Steele, et al., Prion protein (PrPc) positively regulates neural precursor proliferation during developmental and adult mammalian neurogenesis, Proc. Natl Acad. Sci. USA 103 (9) (2006) 3416–3421. [239] A. Abdipranoto, et al., The role of neurogenesis in neurodegenerative diseases and its implications for therapeutic development, CNS Neurol. Disord. Drug Targets 7 (2) (2008) 187–210. [240] B. Waldau, A.K. Shetty, Behavior of neural stem cells in the Alzheimer brain, Cell. Mol. Life Sci. 65 (15) (2008) 2372–2384. [241] J.E. Montgomery, et al., A novel model of retinal ablation demonstrates that the extent of rod cell death regulates the origin of the regenerated zebrafish rod photoreceptors, J. Comp. Neurol. 518 (6) (2010) 800–814. Biochimica et Biophysica Acta 1812 (2011) 381–389 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Behavioural assessments of neurotoxic effects and neurodegeneration in zebrafish☆ Keith B. Tierney ⁎ Department of Biological Sciences, University of Alberta, Z 718 Biological Sciences Building, Edmonton, Alberta, Canada T6G 2E9 a r t i c l e i n f o Article history: Received 2 December 2009 Received in revised form 27 September 2010 Accepted 21 October 2010 Available online 28 October 2010 Keywords: Zebrafish Behaviour Activity Sensorimotor Learning endpoints a b s t r a c t Altered neurological function will generally be behaviourally apparent. Many of the behavioural models pioneered in mammalian models are portable to zebrafish. Tests are available to capture alterations in basic motor function, changes associated with exteroceptive and interoceptive sensory cues, and alterations in learning and memory performance. Excepting some endpoints involving learning, behavioural tests can be carried out at 4 days post fertilization. Given larvae can be reared quickly and in large numbers, and that software solutions are readily available from multiple vendors to automatically test behavioural responses in 96 larvae simultaneously, zebrafish are a potent and rapid model for screening neurological impairments. Coupling current and emerging behavioural endpoints with molecular techniques will permit and accelerate the determination of the mechanisms behind neurotoxicity and degeneration, as well as provide numerous means to test remedial drugs and other therapies. The emphasis of this review is to highlight unexplored/ underutilized behavioural assays for future studies. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Introduction 1.1. The importance of behavioural endpoints In animals, and arguably all life throughout the biological kingdoms, behaviours are ways to integrate into the biotic and abiotic environment. Behaviours are the actions and reactions taken to internal and external cues that are intended to place organisms in a preferable position with respect to fitness. Conditions and situations that cause deviant or impaired behaviours therefore have clear survival implications. The inclusion of a behavioural assessment endpoint in studies of disease and neurotoxin exposure, especially those involving non-lethal impairments, will help address whether sub-organismal changes will affect survival. However, there is another reason to include behavioural endpoints—they can be used to rapidly and effectively detect changes of molecular to system level origins. In vertebrates, all behaviours are achieved through nervous control. However, not all behavioural modifications are of neurological origin. Non-neurological alterations can also affect behaviour— some musculoskeletal issues, for example, such as those arising Abbreviations: AChE, acetylcholinesterase; Anti-AChE, anticholinesterase; CPP, conditioned place preference; CS, conditioned stimulus; dpf, days post fertilization; OKR, optokinetic response; OMR, optomotor response; OP, organophosphorus agent; OSN, olfactory sensory neuron; PEA, phenylethyl alcohol; PTZ, pentylenetetrazole; UCS, unconditioned stimulus ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Tel.: +1 780 492 5339. E-mail address: [email protected]. 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.10.011 through injury and developmental abnormalities, can cause altered behaviours. Malformed limbs or other morphologically-based conditions may be associated with behaviours not apparent in normal individuals. There are also a suite of developmental issues that are related to neurological function, however, they are not related to neurotoxins or neurodegeneration. Cerebral palsy, for instance, can affect gross and/or fine motor control and have very apparent behavioural abnormalities [1]. Additionally, in healthy animals, internal, physiological processes can also modify neuron function and these can result in behaviour phenotypes. For the purposes of this review, only behavioural alterations arising though neurological pathways and directly involving neurotoxin exposure or progressive neurological pathologies will be given consideration. Behavioural alterations arising from neurotoxin exposure and neurodegeneration come about through very different means, even though the behavioural “phenotype” (observable manifestation) may appear similar in some instances. Neurotoxic agents affect the functionality of “normal” neurons and may be reversible and of limited duration. In contrast, neurodegeneration collectively refers to a group of typically irreversible processes that work at the genetic to system level, all of which result in the loss of neurons and/or their functionality. Mechanistically, neurotoxic agents act in one of two ways: in general, non-specific ways (e.g. polar or non-polar narcosis [2]); or through affecting genetic and/or protein receptors. Either mode of action can cause changes at molecular, cellular, system and levels beyond. It must be noted that the first mode of action could include disruption of cells in addition to neurons. With neurodegeneration, mechanisms of action include protein misfolding and conformational disorders (proteopathies) and/or astrogliosis, i.e. an 382 K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 increase in astrocytes due to neuron death. For each mechanism of neuron (or neural tissue) impairment, it should be possible to isolate an individual behavioural phenotype for screening purposes. All behaviours involve motion, and so given the appropriate methods, all are quantifiable. The evolution of software solutions in the last decade or so, including solutions from EthoVision (www.noldus.com), LocoScan (www.cleversysinc.com), Loligo (www.loligosystems.com), and VideoTrack (www.viewpoint.fr) has greatly facilitated the accurate assessment of subtle and directed movements. However, unlike the movement of mammalian models, fish move in three dimensions. In studies of fish motion, movement in the vertical plane is typically ignored or minimized (i.e. by keeping the tank water shallow) [3–5], however, in at least one case, a proprietary protocol was developed to include it [6]. Nevertheless, with a model such as a zebrafish, the best solution may be to minimize water depth, to perhaps twice body length. This depth does not appear to result in overt stress and most fish settle down (acclimate) within 30 min (personal observation). Current technology permits the rapid, and potentially repeated, non-lethal testing of numerous fish (e.g. 4+ adults to 96 larvae) [7,8]. Additionally, a method has been constructed to calculate very subtle differences in the swimming mechanics of individual fish [4]. Caveats exist in behavioural assessments. Organisms are designed with systems that provide resilience and can act to retain higher level responses [9]. Also, factors that affect responses at lower organisational levels, e.g. protein level, may not necessarily have an obvious behavioural phenotype. Behavioural studies clearly need a mechanistically-based rationale for including a motion-based endpoint. 1.2. The relevance of zebrafish models Ten years ago, zebrafish were identified as an up and coming model for genetic disorders and developmental biology in humans [10]. Since this time, the zebrafish genome has been fully sequenced and many genes of high mammalian homology have been identified. Five years ago, zebrafish were proposed as an untapped behaviour-genetic model organism [11]. There are now high-throughput behavioural zebrafishbased screens able to detect specific neurological alterations/pathologies, some of which take advantage of protocols similar in function and purpose to those used on rodents (e.g. olfactory-based endpoints). Furthermore, zebrafish are more amenable to manipulation and behavioural testing during early development. Their transparent larvae facilitate localization of proteins and the use of morpholino manipulation. For these reasons, and given that rodents are more expensive and require greater growth periods than zebrafish, the adoption of zebrafish models continues in typically rodent dominated fields. This review is intended to communicate the current use of behavioural endpoints in zebrafish, and how protocols used with other animals may be adapted. Numerous behavioural assessment avenues remain unexplored. 2. Types of behavioural endpoints As with most animals, fishes exhibit intricate behaviours that depend on locomotion and may be the result of complex decisions [12]. Behavioural responses must be viewed in a three-part hierarchical manner: (1) basic motor responses, which underlie (2) sensorimotor responses, that facilitate and/or integrate with (3) learning and memory (Fig. 1). Although the first “behavioural level,” arguably does not involve true behaviour since it does not necessarily include sensory responses, many studies have included locomotory (activity) changes as a behavioural endpoint [13–15]. While this endpoint may appear simplistic, it does not mean it is not meaningful, in fact, quite the opposite: all behaviours are predicated on the ability to move. Furthermore, in tests of higher-level behaviours, such as those involving learning, it may be difficult to rule out potential neurological issues of the lower founding “levels.” For example, if feeding response decreased in fish exposed to a dissolved neurotoxin such as an organophosphorus (OP) insecticide, the effect could be the result of impaired locomotory Fig. 1. Behavioural responses can be evoked by internal and external signals and/or be altered by impaired neurological condition. With zebrafish, assessment methodologies exist to test neurological performance in absence of signals through to the ability of the brain to retain and integrate sensory signals of diverse origins, which makes zebrafish a powerful model for testing compounds that affect neurons over brief to life-spanning timeframes. CPP = conditioned place preference. ability, and/or impaired olfactory sensory neuron (OSN) function, and/ or cognitive ability. Implicit in neurodegeneration and neurotoxin exposure is that organism condition, and so likely behavioural responses, will change with time. With neurotoxins, there are four temporal phases to describe their effects: direct, secondary, tertiary and quaternary effects. The first two phases deal with the presence and actions of an agent within the organism, and the latter two, without. Direct effects are those local or systemic effects that occur over the short run and would typically be the “intended” effects, such as the actions of an anticholinesterase (anti-AChE) drug. Secondary effects consist of adaptations made over longer periods to restore equilibrium from drug actions, such as upregulation of acetylcholinesterase (AChE) expression. Tertiary effects are those that follow from the discontinuation of administration of an agent that has become physically depended, i.e. involve withdrawal and stress. Over extended periods, tertiary effects may give way to quaternary effects, such as malnutrition. The majority of behavioural assays focus on direct and secondary effects, however, there are studies of tertiary effects, e.g. withdrawal from cocaine [16]. In the following, a diverse array of endpoints using apparatus from simple, one chambered tests, to complex multi-chamber, decisionbased challenges are discussed (Table 1). The majority of the behavioural endpoints can be conducted on zebrafish 3–4 days post fertilization [17,18], excepting learning/memory-based endpoints. 2.1. Basic motor response endpoints Compounds that modulate neuron firing rate have potential to affect locomotory activity. In fishes, locomotory activity endpoints include swimming speed [19], distance covered [14], and turning rate (angular velocity) [20]. Changes in zebrafish activity have been noted in studies investigating the mechanisms of addiction to amphetamines [21], cocaine [16], ethanol [11,22], and nicotine [23], as well as the effects of pesticides [24,25]. Whether any given drug/neurotoxin increases or decreases activity is related to concentration/dose and time. For example, exposure to a low concentrations of D-amphetamine caused hyperactivity while higher concentrations caused hypoactivity [21]. Cholinesterase inhibiting drugs/pesticides are well known to have similar effects [26–29]. Clearly, agonizing stimulatory receptors or inhibiting neurotransmitter degradation will potentially lead to activity increases in the short run that are not sustainable in the long. In fact, persistent stimulation may result in excitotoxicity and neuron apoptosis. Behavioural manifestations of neurotoxin exposure may not appear until later in life. Additionally, some activity changes K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 383 Table 1 Behavioural test endpoints and their associated stimuli available to fishes, including zebrafish. Some endpoints that are underutilized or hold promise in toxicity or degenerative studies are in bold. Apparatus Stimulus Stim. type Type of test References Open arena None Space Food, dead Food, alive Tap Glass bead Predator Predator scent Simulated conspecific Water flow Social Simulated object None Vis Vis; Olf Vis; Olf Aud Aud; Vis Vis Olf Social Motion Vis; Motion Vis Vis Vis; Olf Olf Vis Chemo Drug Vis Pain Vis Vis; Stress Vis Activity levelA Searching behaviourA Appetitive behaviourA Prey captureA Startle responseA Startle responseA Alarm responseA Alarm responseB AggressivenessB UcritB SchoolingB OMRA OKRA Attraction; avoidanceB AttractionB CPPB CPPB CPPB CPPB Aversive CPAB CPPB CPP, CPAB SchoolingB [14,16,20–22,24,25,30] [22,23,70,90] [37] [40,42,43] [7] [58] [39,40] [68–70] [11,22] [83,84] [58,59] [47,49,50] [51,52] [15,74,75] [76–79] [93] [97] [97,98] [22] [16] [99] [100] [22] Confined tube, or flume Moving background (striped drum) Counter-current olfactometer Y-maze T-maze Two-chamber tank Three-chamber, gated Three-chamber, separated Odours⁎ Odours⁎ Preferred location Food Location Shade Electric field Colour Confinement Social Abbreviations: Aud, auditory; CPA, conditioned place aversion; CPP, conditioned place preference; OKR, optokinetic response; Olf, olfactory; OMR, optomotor response; Ucrit, critical swimming speed; Vis, visual. Note: Zebrafish age may affect whether specific endpoints are appropriate, as the endpoints may depend on the development of swimming ability, sensory responses, and memory. In general, assays requiring swimming or sensory-evoked swimming are available as soon as 48 post hatch (~4 dpf; endpoints marked A) [37,106]. Swimming intensive and memory based endpoints have mostly been conducted on adults (approx. ≥3 months), however many of these assays could likely be conducted on juveniles (endpoints marked B). An exception: OMR relies on immobilizing fish, and so fish cannot be N 7 dpf [51]. A caveat: olfactory-based alarm responses may be most robust and least variable at 50 dpf [73]. ⁎ Odours include food, conspecific, pheromones, predator, and synthetic compounds. may only be evident after the removal of the drug, i.e. during withdrawal [16]. Zebrafish activity has proven a viable endpoint for detecting neurological impairments received during development. For example, embryonic zebrafish exposed to chlorpyrifos, an organophosphorus insecticide and ubiquitous environmental pollutant, had reduced swimming activity 6+ days later [24]. If persisting neurological impairment or neurodegeneration are not readily apparent, a drug can be used to evoke activity. Altered neurology will be apparent in a lower response threshold to the drug. The convulsant (seizure inducing drug) pentylenetetrazole (PTZ) has been used in recent studies in zebrafish [30], including high throughput 24 and 96 well plate assays [3,31,32], for just such a purpose. Specifically, zebrafish embryos injected with domoic acid (DA), a neurotoxin of diatom origin, reduced the concentration of PTZ required to evoke stage I and II seizure activity [33]. Activity endpoints can also be used to detect neurodegeneration. Specifically, to model the effects of Parkinson disease, neurotoxic drugs can be injected directly into specific brain regions (using microinjection), although this has historically not carried out on fish [34]. A recent study demonstrated that unilateral injection of the drugs 1,2,3,6-tetrahydropyridine (MPTP) and 6hydroxydopamine (6-OHDA), two neurotoxins with specific targets, reduced the speed and distance zebrafish travelled [14]. Microinjection techniques make the zebrafish model portable to a variety of neural ablation studies. A consideration for activity-based sensory endpoints is that they may not actually be stimulus-free. When a fish or other animal is introduced into a new environment, such as a test tank, they will likely begin searching and forming a cognitive map of their new surroundings [35]. This process will draw upon one or more senses and involve synaptic plasticity (discussed below). Even with elimination of searching behaviour, perhaps accomplished through tank acclimation, activity changes may be based on input from the interoceptive senses, i.e. the perception of internal movement and/ or pain (aspects of the somatosensory system). Activity-based endpoints clearly have application, but they may be more of an umbrella for other neurological impairments isolatable through sense-specific tests. 2.2. Sensorimotor endpoints Fishes have a suite of exteroceptive senses analogous to those of mammals. These include vision (photoreceptor-based), audition (mechanoreceptor-), olfaction, gustatory (both chemosensor-), balance and somatosensory, which includes touch (mechanoreceptor-), pressure, temperature (thermoreceptor-) and pain (nociceptor-). Fishes also posses senses that do not have mammalian analogs, e.g. magnetoreception. Furthermore, some of the analogous senses, e.g. gustation, are morphologically dissimilar—fishes, including zebrafish, have solitary chemosensory cells (SCCs), which are externally-located taste buds [36,37]. And unlike mammals, sound and motion are not only perceived by the ear and vestibular (inertial detection) system, but by an additional system, the lateral line. In this system, neuromast cells structurally similar to inner ear mechanoreceptor “hair” cells, detect changes in water pressure and motion relative to themselves. These cells can also detect sounds and provide information regarding acceleration and velocity [38]. In the below, known and potential sensorimotor endpoints will be discussed. The senses enable reflexive, innate responses to “unconditioned stimuli” (UCS), as well as learned responses to “conditioned stimuli” (CS). Both UCS and CS can be used to test sensorimotor responses, and can be positive (attractive) or aversive. UCS evoked behavioural responses stimuli have been included in this section, but their modification is included in the proceeding section on learning and memory. Positive UCS or CS stimuli may arrive via various senses but uniformly enable the organism to place itself in an actual or perceived improved fitness position that otherwise offers some benefit or reward; example stimuli include food, cover, mate call or odour. Aversive stimuli do the opposite; examples stimuli include electric shocks, abrupt and severe changes in lighting, temperature, sound, 384 K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 (unpleasant) odours, or other sensory assaults causing discomfort. Stimuli can also be neutral, i.e. perceptible but without an associated behaviour [8]. For behavioural endpoints, both CS and UCS stimuli are routinely used in behavioural studies in fishes, including those with and without conditioning. Changes in sensory responses can be due to impaired sensory perception or due to impaired motor performance. Separating perceptual vs. motor impairment in sensory studies may involve using an additional endpoint, such as cell or tissue level assessment of neuron function. Another challenge is that most behaviours are multisensory, and isolating specific impairments may be difficult or impossible. Nevertheless, behavioural endpoints exist to evoke responses through each of the senses. 2.2.1. Visual endpoints Unless experiments are conducted under darkness or with visually occluded fish, all behavioural endpoints arguably have some visual component. In fish, vision specific endpoints include predator (simulated or actual) avoidance, prey capture, optomotor and optokinetic responses. All of these methods have clearly defined endpoints and involve coordination, except the last, which uses immobilized fish. One of the most basic assays draws upon phototaxis, i.e. light seeking behaviour. Zebrafish are active during the day, and so should spend time in light unless they perceive a risky (predation-prone) condition. When given a choice between a lighted or shaded area in an open arena, zebrafish did indeed spend less time under shade (~40%) [22]. Ethanol exposure caused a concentration-dependent decline in the proportion of time spent under shade. With exposures to ethanol, zebrafish sheltering was reduced to ~ 10–15%, i.e. exposure evoked risky behaviour. Such a robust and easily quantifiable behavioural contrast could be exploited to test other neurological impairments. Alarm (antipredator) behaviours are often used in fish research, although they are usually investigated using olfactory cues. Nevertheless, visual cues can be used to evoke alarm [39,40]. Specifically, presenting fish with a large shape modelled after a predator or an actual predatory fish, such as in a neighbouring tank, can evoke dashing, freezing and shelter seeking. Fast start responses (s-starts) have been shown to increase in response to the presentation of a simulated predator (painted falcon tube), and be sensitive to the effects of ethanol exposure [22]. A recent paper suggested that alarm response could be a very valuable tool for mechanistically understanding the roles of signalling molecules, specific receptors and neurons in neural circuitry [41]. One of the advantages to aversive responses is that they are easy to score either manually or through video analysis. However, in some stocks of lab reared fish, innate responses may have been lost. In this case, conditioning would need to be carried out. Visually guided prey capture (as opposed to chemosensory-guided) has been used in a variety of fish species, although not typically on such a small one. Nevertheless, visually guided appetitive behaviours can be carried out even in 7-day-old zebrafish using paramecia [42]. Other prey items such as daphnia, hyalella and brine shrimp have been used in other species, and there is no reason why they cannot also be used with zebrafish. Assay endpoints include time to initiate attack (latency), number of attacked or captured prey, capture efficiency, and time to cessation [40,42,43]. For example, a recent study on striped bass noted that time to prey capture was increased 6 days following exposure to an anti-AChE pesticide, the OP diazinon [44]. Another used zebrafish larvae with laser ablated premotor neurons in the retinotectum and reticulospinal neurons to isolate neural circuitry associated with visually directed prey capture [42]. Tracking software currently uses contrast (i.e. a dark object, the animal, against a light background) to track as many as ten objects in one open arena (EthoVision; www.noldus.com). However, prey may not be easily discernable, and their removal from an arena may not yet be easily quantifiable. A high throughput version of this assays remains for development. A test of optomotor response (OMR) was pioneered over 80 years ago (see [45]) and has been frequently used on a variety of animals since, including mice [46] and zebrafish [47]. The test is one of tracking: animals reflexively keep pace with a rotating a striped or “grated” drum. At a high enough rotation speed, the lines will blur and fish will cease to follow. The parameter of interest is the point at which the animal ceases to track. Excitatory and inhibitory drugs should increase and decrease this point, respectively, and neurodegeneration will obviously affect it. Another way to conduct the test is to hold the drum speed constant but vary the illumination [48]. This method was capable of resolving optomotor differences between strains. A different version of the test replaces the uniform stripes with one large stripe. The large stripe represents a threatening object, and so fish will swim away from it. This test proved capable of detecting retinal degeneration in a “night blind” mutant [49]. A variation of the OMR assay includes using competing visual grating [50]. Specifically, a background reference grating is kept at a constant speed in one direction while a test grating of variable speed is moved in the other. Fish will move in the direction of the more strongly perceived cue. Three scenarios are possible: the fish will track the reference cue when the test cue is of lower contrast, or they will track the test cue when it is of higher contrast, or when there is insufficient contrast between the two, they will stop moving. The null motion point provides an index of visual guided response, with acuity inversely related to contrast. This test identified performance differences from a mutation in a vesicular glutamate transporter. The OMR has been adapted to isolate eye tracking ability in static larval zebrafish [51,52]. In this “optokinetic response” (OKR) assay, larvae are immobilized in methylcellulose gel, placed in the same visual scenario as adults (a circular arena with a rotating striped perimeter), and smooth tracking and rapid saccades with stripe sweep are recorded. Impaired altered visual performance will be evident in decreased saccadic movement [52–54]. The output of eye velocity in degrees/s can be determined using computer software (a protocol is available [55]). A consideration is that fish cannot be N 7 dpf because gel immobilization becomes problematic [51]. The OKR test is functionally equivalent to the finger tracking-based field sobriety test, where the degree at which eye motion switches from smooth tracking to saccadic movement is indicative of intoxication. Zebrafish are a schooling species that exhibits territoriality [56,57]. Although schooling and aggressive behaviour are not solely driven by vision–they can rely on sensory input from the octavolateralis (sound/ motion), etc.–they are predicated on it. The propensity of a small group of zebrafish that was physically but not visually isolated from a larger school to swim near the larger group, i.e. their “social preference,” was inhibited in a concentration-specific manner by ethanol [22]. School cohesion, apparent in “nearest neighbour distance” (NND), may also be used as a sensitive indicator of intoxication [58] and genetic-based differences [59]. Aggressive behaviour (agonistic behaviour) consists of alternating and/or coincident fin displays and attack behaviours. Displays consist of erection of dorsal, pectoral or anal fins, and/or slapping of the caudal fin; attack behaviours include biting motions and directed swimming at the attacker [22]. An inclined mirror can be used to simulate a conspecific, and assess aggression [11,22]. In an “aggressiveness test,” ethanol exposure was correlated with increased time near the mirror and aggressive displays [22]. Automatic video analysis of fin motions obviously poses a challenge, but changes in speed and distance near the mirror would be easy to determine. 2.2.2. Olfactory endpoints In fish, olfaction is so important to so many behaviours that it cannot be done without (reviewed in [60]). This is the reason numerous studies have focused on its impairment through exposure to a variety of neurotoxic contaminants (reviewed in [61]). In humans, olfaction is not so indispensable, however, its loss does correlate with neurodegenerative diseases such as Parkinson's and Alzheimer's [62]. The use of K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 zebrafish olfactory-based behavioural endpoints of neurodegeneration remains largely unexplored. A recent publication suggests olfactory endpoints may also be viable in prion research [63]. In fishes, odorants and odorant classes are associated with specific behavioural responses, which facilitate isolating specific neuronal impairments. Odorants include amino acids, bile salts and pheromones, which evoke feeding and antipredator, social (assumed), alarm and mating-based responses, respectively [60]. Amino acids associated with food, such as L-alanine [64], evoked appetitive behaviours such attraction (directed movement up a concentration gradient) [65]. A mixture of ten amino acids, including L-alanine, evoked increases in turning rate and activity level at 4 dpf [37], suggesting the possibility of using amino acid responses in high throughput assays. Bile salts remain untested as behavioural modulators in zebrafish, but given their efficacy at evoking OSN responses [66], they may be of use in olfactory-based assays. Pheromone-based behavioural endpoints of mating have been restricted to other species (e.g. F-type prostaglandins evoke substrate probing in salmonids [67]), however, the behaviours evoked by alarm pheromone are well studied. Alarm pheromone is anxiogenic and makes for a very reliable endpoint since it is an innate reflex. Alarm responses involve a set of stereotypical behaviours, which include dashing, freezing, jumping and erratic movements [68–70]. All the actions are “fight or flight” in nature, and are associated with increases in stress hormone concentrations [70]. Many fish studies have included alarm response as an endpoint of olfactory neuron function [61]. Numerous neurotoxic agents, including anti-AChE pesticides, alter alarm response, possibly through impairing both peripheral (OSNs) and central neurons [61,71]. Alarm behaviour endpoints are highly sensitive, with some OPs impairing responses with exposures to just 1 ppb [28,71]. Studies wishing to test the functionality of the hypothalamic-pituitary-interrenal (HPI) axis (analogous to the hypothalamic-pituitary-adrenal (HPA) axis in mammals) may benefit from the inclusion of an alarm assay [41]. A recent study found that hypoxanthine 3-N-oxide (H3NO) evoked alarm response in zebrafish [72]. Whether this compound is alarm pheromone or a component thereof, this discovery has great potential to improve the standardization of alarm response testing, since a molecularly based behavioural concentration-response curve can now be constructed. The previous alarm response testing methods involved the use of a skin extract or filtrate derived from the skin of conspecifics. These methods suffer by not being able to standardize the unknown(s) pheromone, however some did endeavour to do so by measuring total protein [6]. In zebrafish, the alarm response may be most robust and least variable 50 dpf [73]. One of the simplest tests for olfactory-evoked motion involves the use of an avoidance trough or “counter current olfactometer.” In this device, water/odour sources flow from either end of rectangular trough and exit in the center without mixing appreciably. In this way, there are effectively two odour zones in the tank, and fish can choose to spend time in either one. As long as a fish is swimming back and forth prior to odorant introduction (i.e. has not preferred an end before stimulus introduction), both attractive and aversive responses can be determined [15,74,75]. This test is not unlike a Y-maze, where odours flow down two arms to a null or mixing zone. Y-maze can be problematic in some cases because fish may elect to make no decision, i.e. not explore the arms. Nevertheless, both Y-mazes and avoidance troughs have been useful in fish studies of the attractive or aversive qualities of amino acids, pheromones and neurotoxic contaminants [76–79]. 2.2.3. Auditory endpoints Even though the anatomy of the zebrafish ear is dissimilar to that of mammals (i.e. no cochlea), the neuron hair cell structure is highly conserved. Additionally, the lateral line cells are structurally and functionally the same as ear hair cells. Both ear and later line cells can be disturbed by compounds of human origin (i.e. “ototoxic” agents). 385 For example, the antibiotic streptomycin can negatively affect fish behaviour [80]. For these reasons, hair cell alteration has mammalian parallel (reviewed in [17]). An auditory-evoked startle response has recently been adapted from rodents to zebrafish [7]. In the study, adult zebrafish were placed individually in an eight tank array and video was recorded from overhead. The stimulus consisted of an automated, mechanical “tap” to the tank bottom. Fish swimming distance was determined over the following 5 s, and the taps were repeated every minute for a total of ten times. This session was then subsequently repeated. This design is particularly interesting because its neurological assessment is twodimensional: it tests an innate, behavioural reflex (which includes integration of sound and space, and depends on physiological condition), as well as the ability to adapt (habituate) rapidly to a stimulus. In their study, they found that both endpoints could show changes due to larval exposure to chlorpyrifos. However, it was arguably the ability to adapt that highlighted the biggest neurological alteration. The difference in startle response magnitude was greater and more significant with repetition. The purpose of this high throughput auditory-based test was to assess the persisting effects of chlorpyrifos exposure, however there is no reason this assay could not be used to screen/assess a variety of neurological conditions/ pathologies. Furthermore, given their findings regarding habituation, studies including other stimuli (such as odorants), could benefit from including such a parameter. 2.2.4. Interoceptive sensory endpoints Fish rely on balance and experience pain. However, unlike mammals, equilibrioception (balance) is accomplished through the octavolateralis system, which includes neurons that perceive sound and motion. Additionally, fish do not rely upon their “limbs” (i.e. fins) for balance. In fact, removing two or more fins has frequently been used as a non-invasive, low impact means of identifying fish [81,82]. Nevertheless, neurotoxins affecting the vestibular portion (i.e. inertial sense) of the inner ear, or neurodegeneration in general, may conceivably affect balance, and so may impair rheotactic ability. Performance deficiencies could be gauged through measuring the time to current orientation. Decades of research on fish swimming performance has relied on mild electrical shock to evoke swimming [83,84]. Detecting altered neurological performance through nociceptor-based response is not common in fish, but it does show promise, particularly with agents/ conditions which modify anxiety. A surprising result of a recent study was that cocaine did not alter a behavioural avoidance of a mild electrical field [16]. Given that establishing a local electric field in an open arena is not difficult, and that fish will instinctively avoid it, this method has promise for screening neurotoxins with anaesthetic effects. 2.3. Learning and memory based endpoints In mammals, the hippocampus is involved with cognitive plasticity. Zebrafish lack a hippocampus, but they do have a developmentally similar region, the lateral pallium, that appears functionally equivalent [85]. Reference memory (i.e. acquired memories) and working memory (i.e. the attentional component of shortterm memory), and their interplay (i.e. cognitive plasticity) can be tested in a variety of ways, including both classical (Pavlovian) and operant (instrumental) conditioning. Both conditioning methods evoke response(s) using a pre-existing behaviour (evoked by either CS or UCS), but for classical conditioning, a neutral stimulus is converted into a positive CS (through their repeated pairing), and in operant conditioning consequences are used to reinforce or extinguish a CS. In the below, endpoints are discussed for (1) attention, for (2) searching (map building) behaviour, for (3) the acquisition of responses to neutral stimuli (classical conditioning), as well as for (4) the 386 K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 reinforcement or extinguishment of positive and negative stimuli (operant conditioning). 2.3.1. Attentional endpoints Behaviours generally result from the integration of multiple signals. What is important is the ability to discern important signals amidst background noise. Neurological filtering in vertebrates has been referred to as the “cocktail party effect,” coined to describe the ability of a person to hear a signal such as their name spoken in a room full of chatting people [86]. Implicit in “selective attention” is that distracting signals be ignored. One assessment methodology used in non-human primates involves coupling a positive stimulus, such as food, with the occurrence of another stimulus, such as a particular visual signal, and varying delay between the two [87]. Distractability can be assessed though inserting irrelevant stimuli in the delay period. Since filtering draws upon reference memory and working memory, it is a useful correlate of neurological function. While numerous paradigms exist to test selective attention-based tests in rodents and primates [87,88], no obvious correlate currently exist for zebrafish. However, a functionally similar test might be carried out by challenging zebrafish with competing cues of fitness relevance. For example, fish could be exposed to a food odorant, which would initiate appetitive behavioural response, then interrupted with an alarm cue, which should evoke anti-predator behaviour. The degree and rate at which the appetitive behaviour is replaced and antipredator responses exhibited (i.e. evokes attention) could prove meaningful endpoints of sensorimotor function and cognition. To explicitly test memory function, the UCS food and/or predator cues could be replaced with CS cues, such as behaviourally neutral amino acids or PEA. An attentional endpoint could be especially useful in testing cholinergic system functionality, since working memory is dependent on it [89], and since many drugs agonize or antagonize cholinergic receptors and neurodegenerative diseases such as Alzheimer's impair it. 2.3.2. Searching (spatial) endpoints Introduction to new surroundings will stimulate searching behaviour and map building [35]. The formation of a new cognitive map is another form of neural plasticity and therefore a possible performance metric for neurological impairment. Numerous studies have already validated that searching behaviour in an open arena is a meaningful endpoint in zebrafish [22,70,90]. Searching behaviour was recently used in combination with two anxiogenic factors, alarm pheromone or caffeine, and two axiolytic factors, ethanol and fluoxetine (i.e. the SSRI Prozac), as an endpoint in “a novel tank test” [70]. Release into a new tank evoked a bottom seeking in control fish, which was followed by subsequent upper tank exploration. With alarm pheromone and caffeine exposure, the number of upper tank explorations was reduced, and the time taken to explore (latency) increased. With exposure to ethanol or fluoxetine, the reverse pattern was noted: transitions to the upper tank increased, and the latency to upper tank searching was greatly diminished. Nicotine had the opposite effect, fish were not as inclined to dive to the tank bottom [23]. Spatial memory is mediated through the lateral pallium. Studies have found that lesioning the functionally equivalent region in rats (hippocampus), abolished searching behaviour [91]. Given the ability to focally lesion larval zebrafish brain tissue and neurons using laser ablation [42], studies of searching hold promise for cognitive endpoints. A tried and true test of spatial memory in rats is the Morris water navigation task, which is essentially map building under stress and with visual cue reference. The rate of learning of the platform location can be gauged by the latency to platform discovery and distance covered in successive trials. Learning under stress in fish could be accomplished in a variety of means, such as by the ability to find an escape portal (to another potion of a test tank), under the threat of predation, evoked by visual or olfactory stimuli. One zebrafish study using a net as a freight stimulus showed that the latency to discover an escape portal decreased significantly over six training sessions [92]. 2.3.3. Classical conditioning endpoints Zebrafish studies have demonstrated that both visual [93] and olfactory [8] cues unassociated with behavioural response(s) can be converted to behavioural response cues. For example, over a few (b10) trials, zebrafish were able to visibly discriminate and select coloured arms of a T-maze where they would receive a food reward [93]. As for odorant-based learning, a recent study demonstrated that by pairing PEA (a neutral stimulus) with food odours (L-alanine and Lvaline) over a number of training sessions, appetitive swimming behaviour (turning rate) could be evoked by the addition of PEA alone (i.e. it was converted to a CS) [8]. Classical conditioning endpoints may be increasingly adopted since conditioning can be achieved rapidly. For example, just one pairing of a neutral stimulus (red light) with alarm pheromone (UCS) evoked significant conditioning in a related species (fathead minnow) [94]. These data suggest that the strength of the innate response determines the rate of learning. In many species of vertebrates, shaping is used to evoke a desired behaviour. This method consists of reinforcing successive steps towards a target behaviour. This method of reference memory creation may be unreasonable for fish; however autoshaping may be a viable option. Autoshaping was developed in pigeons and consisted of the pairing of a visual cue (light) with food. Initially the pigeons will respond (peck) to the UCS, but over time will respond to the CS alone [95]. There is no reason why this model would not also work in fish. Zebrafish will swim into odour sources and sometimes “bite” at water inflows including food odour (personal observation). With repeated trials, they could be expected to bite at a water flow source alone. 2.3.4. Operant conditioning endpoints A variety of methods using open arena to multi-chambered arenas exist to test stimulus reinforcement or extinguishment. Open arenas have been used for tests involving food odour and predator scent. In the first, an appetitive response to food flake was enhanced by pairing it with amino acids food odours [8]. In the second, a study of a closely related species (fathead minnow) found that it took 6 to 8 days and two to four days to visually and olfactorily acquire a predator, presumably by the pairing of a UCS (alarm pheromone) with visual cues and scent, respectively [39]. Visual and odorant cues clearly have potential for assessing cognition, and the differential time course could provide a valuable diagnostic tool. A common method of operant conditioning involves presenting an animal with one or more choices between arena arms or areas, and based on the decision, either rewarding the animal with food or punishing it with pain or stress. A popular paradigm is “conditioned place preference” (CPP) or sometimes just “place preference” (PP), in which animals are rewarded with a drug after selecting a particular location [96]. The reinforcing stimulus needs to be a positive psychoactive stimulant, such as cocaine [97] or amphetamine [98]. A version of CPP is “conditioned place avoidance” (CPA), which is where the conditioned area was initially avoided [99]. For example, by “addicting” zebrafish to D-amphetamine, preference for a side of the tank with two large freight inducing black dots could be evoked [99]. In general, CPP and CPA are good models for testing the neurological bases for addiction, but they remain largely unadopted but available in toxicity or degenerative studies. At least two methods exist to test vision-based learning in fishes, the T-maze and the three chambered arena. Both methods share some similarities to the radial arm maze (RAM), frequently used in mammalian models. In the T-maze, fish are introduced into the long arm (base of the T) and they will generally swim to one or both of the sort arms. However, as many as 5% fish may not move from the base K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 arm unless provoked [97]. The intended outcome is for a fish to favour one arm based on the desirability of the settings (“spatial preference”) [97], or a reward coupled to a visual cue [93]. In tests using the three chambered arena, fish are introduced to a center area that is flanked on either side by gated chambers [100]. To generate avoidance behaviour of one of the side chamber, fish are penalized (using confinement) for making repeated selections of one tank side (in general, to condition an animal to avoid an area based on a visual cue, nociceptor and stress based methods are used [101]). The three chamber confinement test demonstrated that cholinergic agonization can improve the number of correct decisions [100]. In all operant models discussed above, a conditioned CS will lose its behavioural stimulatory efficacy as its repeated presentation goes unpaired with the initial pre-existing, positive stimulus [8]. As with acquisitional endpoints, the rate of change in the reinforced decision can provide a correlate of neurological function [102]. 3. Problems associated with behavioural endpoints Variation in responses/traits/parameters increases with organisational level, and so behaviours are obviously the most variable of organismal responses. Additionally, in fish, as other animals, their state will affect memory recall and motivation, and they are not without personality. Furthermore, with aquatic species, special consideration must be made for the water source. Municipal water is often rife with neurotoxic contaminants, which may vary in concentration considerably even over brief time periods. Water needs to be carbon filtered and tested to remove/account for any confounding chemical effects. An appreciable amount of individual variation can be accounted for by classifying behavioural phenotypes, such as low or high activity subgroups. Doing so will increase the power of the test; however, not all traits that appear obviously related to behaviour may be so. In juvenile brook trout, for example, swimming performance ability was not associated with actual (conative) swimming activity [103]. Some temporal issues can be exploited; food responsiveness and willingness to take risks, for example, are enhanced with brief periods of food deprivation [65,103]. Social status can also affect stress level which will in turn modulate activity and stimuli responses [103]. If social endpoint was to be used, accounting for dominance behaviour in fish may present a challenge. Careful observation and selective removal of individuals of stronger resource utilization (e.g. food consumption) could help resolve variation. 4. Conclusions Vertebrates have been proposed as the best models for elucidating mechanisms of neurodegeneration [104]. Among vertebrate models, zebrafish are finding ever increasing application. In studies of drugs, zebrafish make for excellent models, since delivery can be achieved in many cases by simple addition to the tank water. Furthermore, microinjection techniques are now available to deliver agents directly to specific cells/tissues, even within larval fish [105]. This review has highlighted behavioural endpoints available to zebrafish that probe basic neurological function, that test behaviours predicated on the function of diverse types of sensory neurons, and that investigate cognitive performance. Some behavioural endpoints test multiple neurologically-based abilities; this review is intended to provide a template by which to link and/or apportion specific physiological, perceptual and cognitive impairments. Most methodologies described herein have very tight parallel with mammalian models. Furthermore, many behavioural tests, especially those using high throughput screens, remain in evolution. The future holds promise for the development of a suite of standardized behavioural assays that can be used to determine the mechanisms by which neurotoxic effects and neurodegenerative diseases develop and progress. 387 Acknowledgments The author is grateful to W. Ted Allison for motivation in doing this work, and for grants to KBT from the Canadian Wildlife Federation (CWF) and the University of Alberta. References [1] R.N. Boyd, M.E. Morris, H.K. Graham, Management of upper limb dysfunction in children with cerebral palsy: a systematic review, Eur. J. Neurol. 8 (2001) 150–166. [2] D.W. Roberts, J.F. Costello, Mechanisms of action for general and polar narcosis: a difference in dimension, QSAR Comb. Sci. 22 (2003) 226–233. [3] S.C. Baraban, M.R. Taylor, P.A. Castro, H. Baier, Pentylenetetrazole induced changes in zebrafish behavior, neural activity and c-fos expression, Neuroscience 131 (2005) 759–768. [4] E. Fontaine, D. Lentink, S. Kranenbarg, U.K. Muller, J.L. van Leeuwen, A.H. Barr, J.W. Burdick, Automated visual tracking for studying the ontogeny of zebrafish swimming, J. Exp. Biol. 211 (2008) 1305–1316. [5] C.S. Lam, V. Korzh, U. Strahle, Zebrafish embryos are susceptible to the dopaminergic neurotoxin MPTP, Eur. J. Neurosci. 21 (2005) 1758–1762. [6] J.F. Sandahl, D.H. Baldwin, J.J. Jenkins, N.L. Scholz, A sensory system at the interface between urban stormwater runoff and salmon survival, Environ. Sci. Technol. 41 (2007) 2998–3004. [7] D. Eddins, D. Cerutti, P. Williams, E. Linney, E.D. Levin, Zebrafish provide a sensitive model of persisting neurobehavioral effects of developmental chlorpyrifos exposure: Comparison with nicotine and pilocarpine effects and relationship to dopamine deficits, Neurotoxicology and Teratology In Press, Corrected Proof (2009). [8] O.R. Braubach, H.-D. Wood, S. Gadbois, A. Fine, R.P. Croll, Olfactory conditioning in the zebrafish (Danio rerio), Behav. Brain Res. 198 (2009) 190–198. [9] K.B. Tierney, C.J. Kennedy, Chapter 6: Background Toxicology, in: P.J. Walsh, S.L. Smith, L.E. Fleming, H. Solo-Gabriele, W.H. Gerwick (Eds.), Oceans and Human Health: Risks and Remedies from the Seas, vol. 1, Elsevier Science Publishers, New York, 2008, pp. 101–120. [10] L.I. Zon, Zebrafish: A new model for human disease, Genome Res. 9 (1999) 99–100. [11] R. Gerlai, Zebra fish: an uncharted behavior genetic model, Behav. Genet. 33 (2003) 461–468. [12] V.A. Braithwaite, Cognitive ability in fish, in: K.A. Sloman, R.W. Wilson, S. Balshine (Eds.), Behaviour and physiology of fish, vol. 24, Academic Elsevier Press, San Diego, 2006, pp. 1–37. [13] W. Boehmler, T. Carr, C. Thisse, B. Thisse, V.A. Canfield, R. Levenson, D4 Dopamine receptor genes of zebrafish and effects of the antipsychotic clozapine on larval swimming behaviour, Genes Brain Behav. 6 (2007) 155–166. [14] O.V. Anichtchik, J. Kaslin, N. Peitsaro, M. Scheinin, P. Panula, Neurochemical and behavioural changes in zebrafish Danio rerio after systemic administration of 6-hydroxydopamine and 1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine, J. Neurochem. 88 (2004) 443–453. [15] K.B. Tierney, C.R. Singh, P.S. Ross, C.J. Kennedy, Relating olfactory neurotoxicity to altered olfactory-mediated behaviors in rainbow trout exposed to three currently-used pesticides, Aquat. Toxicol. 81 (2007) 55–64. [16] M.A. López-Patiño, L. Yu, H. Cabral, I.V. Zhdanova, Anxiogenic effects of cocaine withdrawal in zebrafish, Physiol. Behav. 93 (2008) 160–171. [17] T.T. Whitfield, P. Mburu, R.E. Hardisty-Hughes, S.D.M. Brown, Models of congenital deafness: mouse and zebrafish, Drug Discov. Today: Disease Models 2 (2005) 85–92. [18] L. Li, Zebrafish mutants: behavioral genetic studies of visual system defects, Dev. Dyn. 221 (2001) 365–372. [19] S.L. Beauvais, S.B. Jones, J.T. Parris, S.K. Brewer, E.E. Little, Cholinergic and behavioral neurotoxicity of carbaryl and cadmium to larval rainbow trout (Oncorhynchus mykiss), Ecotoxicol. Environ. Saf. 49 (2001) 84–90. [20] S.L. Beauvais, S.B. Jones, S.K. Brewer, E.E. Little, Physiological measures of neurotoxicity of diazinon and malathion to larval rainbow trout (Oncorhynchus mykiss) and their correlation with behavioral measures, Environ. Toxicol. Chem. 19 (2000) 1875–1880. [21] T.D. Irons, R.C. MacPhail, D.L. Hunter, S. Padilla, Acute neuroactive drug exposures alter locomotor activity in larval zebrafish, Neurotoxicol. Teratol. 32 (2010) 84–90. [22] R. Gerlai, M. Lahav, S. Guo, A. Rosenthal, Drinks like a fish: zebra fish (Danio rerio) as a behavior genetic model to study alcohol effects, Pharmacol. Biochem. Behav. 67 (2000) 773–782. [23] E.D. Levin, Z. Bencan, D.T. Cerutti, Anxiolytic effects of nicotine in zebrafish, Physiol. Behav. 90 (2007) 54–58. [24] E.D. Levin, H.A. Swain, S. Donerly, E. Linney, Developmental chlorpyrifos effects on hatchling zebrafish swimming behavior, Neurotoxicol. Teratol. 26 (2004) 719–723. [25] C.E.W. Steinberg, R. Lorenz, O.H. Spieser, Effects of atrazine on swimming behavior of zebrafish, Brachydanio rerio, Water Res. 29 (1995) 981–985. [26] K.N. Baer, K. Olivier, C.N. Pope, Influence of temperature and dissolved oxygen on the acute toxicity of profenofos to fathead minnows (Pimephales promelas), Drug Chem. Toxicol. 25 (2002) 231–245. [27] S. Bretaud, P. Saglio, J.-P. Toutant, Effets du carbofuran sur l'activité de l'acétylcholinestérase cérébrale et sur l'activité de nage chez Carassius auratus (Cyprinidae), Cybium 25 (2001) 33–40. 388 K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 [28] J.F. Sandahl, D.H. Baldwin, J.J. Jenkins, N.L. Scholz, Comparative thresholds for acetylcholinesterase inhibition and behavioral impairment in coho salmon exposed to chlorpyrifos, Environ. Toxicol. Chem. 24 (2005) 136–145. [29] S.K. Brewer, E.E. Little, A.J. DeLonay, S.L. Beauvais, S.B. Jones, M.R. Ellersieck, Behavioral dysfunctions correlate to altered physiology in rainbow trout (Oncorynchus mykiss) exposed to cholinesterase-inhibiting chemicals, Arch. Environ. Contam. Toxicol. 40 (2001) 70–76. [30] K.B. Tierney, X. Ren, Z. Alyasha'e, B. Zielinski, Towards a mechanistic understanding of food odor driven motion using zebrafish (Danio rerio), 15th International Symposium on Olfaction and Taste, vol. 33, Oxford Univ Press, San Francisco, CA, 2008, pp. S156–S157. [31] M.J. Winter, W.S. Redfern, A.J. Hayfield, S.F. Owen, J.-P. Valentin, T.H. Hutchinson, Validation of a larval zebrafish locomotor assay for assessing the seizure liability of early-stage development drugs, Journal of Pharmacological and Toxicological Methods 57, 176–187. [32] S.C. Baraban, M.T. Dinday, P.A. Castro, S. Chege, S. Guyenet, M.R. Taylor, A largescale mutagenesis screen to identify seizure-resistant zebrafish, Epilepsia 48 (2007) 1151–1157. [33] J.A. Tiedeken, J.S. Ramsdell, Embryonic exposure to domoic acid increases the susceptibility of zebrafish larvae to the chemical convulsant pentylenetetrazole, Environ. Health Perspect. 115 (2007) 1547–1552. [34] M.F. Beal, Experimental models of Parkinson's disease, Nat. Rev. Neurosci. 2 (2001) 325–334. [35] W.J. Bell, Searching behaviour: the behavioural ecology of finding resources, Chapman and Hall, Cambridge, 1991. [36] K. Kotrschal, W.-D. Krautgartner, A. Hansen, Ontogeny of the solitary chemosensory cells in the zebrafish, Danio rerio, Chem. Senses 22 (1997) 111–118. [37] S.M. Lindsay, R.G. Vogt, Behavioral responses of newly hatched zebrafish (Danio rerio) to amino acid chemostimulants, Chem. Senses 29 (2004) 93–100. [38] C.B. Braun, S. Coombs, R.R. Fay, What is the nature of multisensory interaction between octavolateralis sub-systems? Brain Behav. Evol. 59 (2002) 162–176. [39] G.E. Brown, D.P. Chivers, R.J.F. Smith, Differential learning rates of chemical versus visual cues of a northern pike by fathead minnows in a natural habitat, Environ. Biol. Fish. 49 (1997) 89–96. [40] V.N. Mikheev, J. Wanzenbock, A.F. Pasternak, Effects of predator-induced visual and olfactory cues on 0+ perch (Perca fluviatilis L.) foraging behaviour, Ecol. Freshw. Fish 15 (2006) 111–117. [41] S.J. Jesuthasan, A.S. Mathuru, The alarm response in zebrafish: Innate fear in a vertebrate genetic model, J. Neurogenet. 22 (2008) 211–228. [42] E. Gahtan, P. Tanger, H. Baier, Visual prey capture in larval zebrafish is controlled by identified reticulospinal neurons downstream of the tectum, J. Neurosci. 25 (2005) 9294–9303. [43] J. Savitz, L. Bardygula-Nonn, Behavioral interactions between coho (Oncorhynchus kisutch) and chinook salmon (Oncorhynchus tshawytscha) and prey fish species, Ecol. Freshw. Fish 6 (1997) 190–195. [44] K.M. Gaworecki, A.P. Roberts, N. Ellis, A.D. Sowers, S.J. Klaine, Biochemical and behavioral effects of diazinon exposure in hybrid striped bass, Environ. Toxicol. Chem. 28 (2009) 105–112. [45] C. Schlieper, Farbensinn der tiere und optomotorische reaktionen, J. Comp. Physiol. A Neuroethol. Sens. Neural Behav. Physiol. 6 (1927) 453–472. [46] J. Abdeljalil, M. Hamid, O. Abdel-mouttalib, R. Stéphane, R. Raymond, A. Johan, S. José, C. Pierre, P. Serge, The optomotor response: a robust first-line visual screening method for mice, Vis. Res. 45 (2005) 1439–1446. [47] A. Krauss, C. Neumeyer, Wavelength dependence of the optomotor response in zebrafish (Danio rerio), Vis. Res. 43 (2003) 1275–1284. [48] K. Ijiri, Vestibular and visual contribution to fish behavior under microgravity, Adv. Space Res. 25 (2000) 1997–2006. [49] L. Li, J.E. Dowling, A dominant form of inherited retinal degeneration caused by a non-photoreceptor cell-specific mutation, Proc. Natl Acad. Sci. USA 94 (1997) 11645–11650. [50] M.C. Smear, H.W. Tao, W. Staub, M.B. Orger, N.J. Gosse, Y. Liu, K. Takahashi, M.-m. Poo, H. Baier, Vesicular glutamate transport at a central synapse limits the acuity of visual perception in zebrafish, Neuron 53 (2007) 65–77. [51] S.E. Brockerhoff, J.B. Hurley, U. Janssen-Bienhold, S.C. Neuhauss, W. Driever, J.E. Dowling, A behavioral screen for isolating zebrafish mutants with visual system defects, Proc. Natl Acad. Sci. USA 92 (1995) 10545–10549. [52] O. Rinner, J.M. Rick, S.C.F. Neuhauss, Contrast sensitivity, spatial and temporal tuning of the larval zebrafish optokinetic response, Investig. Ophthalmol. Vis. Sci. 46 (2005) 137–142. [53] F.M. Richards, W.K. Alderton, G.M. Kimber, Z. Liu, I. Strang, W.S. Redfern, J.P. Valentin, M.J. Winter, T.H. Hutchinson, Validation of the use of zebrafish larvae in visual safety assessment, J. Pharmacol. Toxicol. Meth. 58 (2008) 50–58. [54] T.P. Barros, W.K. Alderton, H.M. Reynolds, A.G. Roach, S. Berghmans, Zebrafish: an emerging technology for in vivo pharmacological assessment to identify potential safety liabilities in early drug discovery, Br. J. Pharmacol. 154 (2008) 1400–1413. [55] C. Hodel, S.C.F. Neuhauss, Computer-based analysis of the optokinetic response in zebrafish larvae, Cold Spring Harbor Protocols 2008 (2008)8 pdb.prot4961-. [56] R. Spence, G. Gerlach, C. Lawrence, C. Smith, The behaviour and ecology of the zebrafish, Danio rerio, Biol. Rev. 83 (2008) 13–34. [57] C. Brown, K.N. Laland, Social learning in fishes: a review, Fish Fish. 4 (2003) 280–288. [58] C.A. Dlugos, R.A. Rabin, Ethanol effects on three strains of zebrafish: model system for genetic investigations, Pharmacol. Biochem. Behav. 74 (2003) 471–480. [59] K.B. Tierney, D.A. Paterson, C.J. Kennedy, The influence of maternal condition on offspring performance in sockeye salmon (Oncorhynchus nerka Walbaum), J. Fish Biol. 75 (2009) 1244–1257. [60] H. Hino, N.G. Miles, H. Bandoh, H. Ueda, Molecular biological research on olfactory chemoreception in fishes, J. Fish Biol. 75 (2009) 945–959. [61] K.B. Tierney, D.H. Baldwin, T.J. Hara, P.S. Ross, N.L. Scholz, C.J. Kennedy, Review: olfactory toxicity in fishes, Aquat. Toxicol. (2009) 1–25, doi:10.1016/j. aquatox.2009.09.019. [62] C. Hawkes, Olfaction in neurodegenerative disorder, Mov. Disord. 18 (2003) 364–372. [63] C.E. Le Pichon, M.T. Valley, M. Polymenidou, A.T. Chesler, B.T. Sagdullaev, A. Aguzzi, S. Firestein, Olfactory behavior and physiology are disrupted in prion protein knockout mice, Nat. Neurosci. 12 (2009) 60–69. [64] W.E.S. Carr, J.C. Netherton III, R.A. Gleeson, C.D. Derby, Stimulants of feeding behavior in fish: analyses of tissues of diverse marine organisms, Biol. Bull. 190 (1996) 149–160. [65] C.W. Steele, D.W. Owens, A.D. Scarfe, Attraction of zebrafish, Brachydanio rerio, to alanine and its suppression by copper, J. Fish Biol. 36 (1990) 341–352. [66] W.C. Michel, D.S. Derbidge, Evidence of distinct amino acid and bile salt receptors in the olfactory system of the zebrafish, Danio rerio, Brain Res. 764 (1997) 179–187. [67] F. Laberge, T.J. Hara, Behavioural and electrophysiological responses to F-prostaglandins, putative spawning pheromones, in three salmonid fishes, J. Fish Biol. 62 (2003) 206–221. [68] K. von Frisch, Über einen Schreckstoff der Fischhaut und seine biologische Bedeutung, Z. Vgl. Physiol. 29 (1941) 46–145. [69] N. Speedie, R. Gerlai, Alarm substance induced behavioral responses in zebrafish (Danio rerio), Behav. Brain Res. 188 (2008) 168–177. [70] R.J. Egan, C.L. Bergner, P.C. Hart, J.M. Cachat, P.R. Canavello, M.F. Elegante, S.I. Elkhayat, B.K. Bartels, A.K. Tien, D.H. Tien, S. Mohnot, E. Beeson, E. Glasgow, H. Amri, Z. Zukowska, A.V. Kalueff, Understanding behavioral and physiological phenotypes of stress and anxiety in zebrafish, Behav. Brain Res. 205 (2009) 38–44. [71] N.L. Scholz, N.K. Truelove, B.L. French, B.A. Berejikian, T.P. Quinn, E. Casillas, T.K. Collier, Diazinon disrupts antipredator and homing behaviors in chinook salmon (Oncorhynchus tshawytscha), Can. J. Fish. Aquat. Sci. 57 (2000) 1911–1918. [72] K.V. Parra, J.C. Adrian Jr., R. Gerlai, The synthetic substance hypoxanthine 3-N-oxide elicits alarm reactions in zebrafish (Danio rerio), Behav. Brain Res. 205 (2009) 336–341. [73] S.R. Blechinger, R.C. Kusch, K. Haugo, C. Matz, D.P. Chivers, P.H. Krone, Brief embryonic cadmium exposure induces a stress response and cell death in the developing olfactory system followed by long-term olfactory deficits in juvenile zebrafish, Toxicol. Appl. Pharmacol. 224 (2007) 72–80. [74] E.E. Little, S.E. Finger, Swimming behavior as an indicator of sublethal toxicity in fish, Environ. Toxicol. Chem. 9 (1990) 13–19. [75] J.A. Hansen, D.F. Woodward, E.E. Little, A.J. Delonay, H.L. Bergman, Behavioral avoidance: possible mechanism for explaining abundance and distribution of trout species in a metal-impacted river, Environ. Toxicol. Chem. 18 (1999) 313–317. [76] Y. Ishida, H. Kobayashi, Avoidance behavior of carp to pesticides and decrease of the avoidance threshold by addition of sodium lauryl sulfate, Fish. Sci. Tokyo 61 (1995) 441–446. [77] J.D. Morgan, G.A. Vigers, A.P. Farrell, D.M. Janz, J.F. Manville, Acute avoidance reactions and behavioral responses of juvenile rainbow trout (Oncorhynchus mykiss) to Garlon 4® Garlon 3A® and Vision® herbicides, Environ. Toxicol. Chem. 10 (1991) 73–79. [78] H. Yambe, M. Shindo, F. Yamazaki, A releaser pheromone that attracts males in the urine of mature female masu salmon, J. Fish Biol. 55 (1999) 158–171. [79] T.J. Hara, Feeding behaviour in some teleosts is triggered by single amino acids primarily through olfaction, J. Fish Biol. 68 (2006) 810–825. [80] J.M. Gardiner, J. Atema, Sharks need the lateral line to locate odor sources: rheotaxis and eddy chemotaxis, J. Exp. Biol. 210 (2007) 1925–1934. [81] I.K. Birtwell, R. Fink, D. Brand, R. Alexander, C.D. McAllister, Survival of pink salmon (Oncorhynchus gorbuscha) fry to adulthood following a 10-day exposure to the aromatic hydrocarbon water-soluble fraction of crude oil and release to the Pacific Ocean, Can. J. Fish. Aquat. Sci. 56 (1999) 2087–2098. [82] K.B. Tierney, E. Stockner, C.J. Kennedy, Changes in immunological characteristics and disease resistance of juvenile coho salmon (Oncorhynchus kisutch Walbaum) in response to dehydroabietic acid exposure under varying thermal conditions, Water Qual. Res. J. Can. 39 (2004) 176–184. [83] C.H. Hammer, Fatigue and exercise tests with fish, Comp. Biochem. Physiol. A Physiol. 112 (1995) 1–20. [84] K.B. Tierney, M. Casselman, S. Takeda, A.P. Farrell, C.J. Kennedy, The relationship between cholinesterase inhibition and two types of swimming performance in chlorpyrifos-exposed coho salmon (Oncorhynchus kisutch), Environ. Toxicol. Chem. 26 (2007) 998–1004. [85] F. Rodríguez, J.C. López, J.P. Vargas, C. Broglio, Y. Gómez, C. Salas, Spatial memory and hippocampal pallium through vertebrate evolution: insights from reptiles and teleost fish, Brain Res. Bull. 57 (2002) 499–503. [86] E.C. Cherry, Some experiments on the recognition of speech, with one and with two ears, J. Acoust. Soc. Am. 25 (1953) 975–979. [87] M.A. Prendergast, Assessment of distractability in non-human primates performing a delayed matching task, in: J.J. Buccafusco (Ed.), Methods of Behavior Analysis in Neuroscience, CRC Press, Boca Raton, 2001, pp. 123–140. [88] P.J. Bushnell, B.J. Strupp, Assessing attention in rodents, in: J.J. Buccafusco (Ed.), Methods of Behavior Analysis in Neuroscience, CRC Press, Boca Raton, 2009. K.B. Tierney / Biochimica et Biophysica Acta 1812 (2011) 381–389 [89] M.I. Miranda, G. Ferreira, L. Ramírez-Lugo, F. Bermúdez-Rattoni, Role of cholinergic system on the construction of memories: taste memory encoding, Neurobiol. Learn. Mem. 80 (2003) 211–222. [90] R. Blaser, R. Gerlai, Behavioral phenotyping in zebrafish: comparison of three behavioral quantification methods, Behav. Res. Meth. 38 (2006) 456–469. [91] J. O'Keefe, L. Nadel, The hippocampus as a cognitive map, Oxford University Press, Oxford, 1978. [92] D. Arthur, E. Levin, Spatial and non-spatial visual discrimination learning in zebrafish (Danio rerio), Anim. Cogn. 4 (2001) 125–131. [93] R.M. Colwill, M.P. Raymond, L. Ferreira, H. Escudero, Visual discrimination learning in zebrafish (Danio rerio), Behav. Process. 70 (2005) 19–31. [94] W.K. Yunker, D.E. Wein, B.D. Wisenden, Conditioned alarm behavior in fathead minnows (Pimephales promelas) resulting from association of chemical alarm pheromone with a nonbiological visual stimulus, J. Chem. Ecol. 25 (1999) 2677–2686. [95] C.M. Locurto, H.S. Terrace, J. Gibbon, Autoshaping and conditioning theory, in: C.M. Locurto, H.S. Terrace, J. Gibbon (Eds.), Autoshaping and Conditioning Theory, Academic Press, New York, 1981, p. 313. [96] J.R. James, R. Young, J.A. Rosecrans, Conditioned place preference: an approach to evaluating positive and negative drug-induced stimuli, in: J.J. Buccafusco (Ed.), Methods of Behavior Analysis in Neuroscience, CRC Press, Boca Raton, 2001, pp. 81–90. [97] T. Darland, J.E. Dowling, Behavioral screening for cocaine sensitivity in mutagenized zebrafish, Proc. Natl Acad. Sci. USA 98 (2001) 11691–11696. 389 [98] J. Ninkovic, A. Folchert, Y.V. Makhankov, S.C.F. Neuhauss, I. Sillaber, U. Straehle, L. Bally-Cuif, Genetic identification of AChE as a positive modulator of addiction to the psychostimulant D-amphetamine in zebrafish, J. Neurobiol. 66 (2006) 463–475. [99] J. Ninkovic, L. Bally-Cuif, The zebrafish as a model system for assessing the reinforcing properties of drugs of abuse, Methods 39 (2006) 262–274. [100] E. Levin, J. Limpuangthip, T. Rachakonda, M. Peterson, Timing of nicotine effects on learning in zebrafish, Psychopharmacology 184 (2006) 547–552. [101] J. Garcia, P.S. Lasiter, F. Bermudez-Rattoni, D.A. Deems, A general theory of aversion learning, Ann. NY Acad. Sci. 443 (1985) 8–21. [102] T.E. Orr, P.A. Walters, R.L. Elkins, Fundamentals, methodologies, and uses of taste aversion learning, in: Buccafusco (Ed.), Methods of Behavior Analysis in Neuroscience, CRC Press, Boca Raton, 2001, pp. 51–70. [103] M. Farwell, R.L. McLaughlin, Alternative foraging tactics and risk taking in brook charr (Salvelinus fontinalis), Behav. Ecol. 20 (2009) 913–921. [104] J.D. Cooper, C. Russell, H.M. Mitchison, Progress towards understanding disease mechanisms in small vertebrate models of neuronal ceroid lipofuscinosis, Biochim Biophys Acta (BBA), Mol. Basis Dis. 1762 (2006) 873–889. [105] C. Cianciolo Cosentino, B.L. Roman, I.A. Drummond, N.A. Hukriede, Intravenous microinjections of zebrafish larvae to study acute kidney injury, J. Vis. Exp. 42 (2010) e2079. [106] J.S.S. Easter, G.N. Nicola, The development of vision in the zebrafish (Danio rerio), Dev. Biol. 180 (1996) 646–663. Biochimica et Biophysica Acta 1812 (2011) 390–401 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Aberrant forebrain signaling during early development underlies the generation of holoprosencephaly and coloboma☆ Patricia A. Gongal a,1, Curtis R. French a, Andrew J. Waskiewicz a,b,c,⁎ a b c Department of Biological Sciences, University of Alberta, Edmonton, Canada T6G 2E9 Centre for Neuroscience, University of Alberta, Edmonton, Canada T6G 2E9 Women & Children's Health Research Institute, University of Alberta, Edmonton, Canada T6G 2E9 a r t i c l e i n f o Article history: Received 1 December 2009 Accepted 8 September 2010 Available online 16 September 2010 Keywords: Holoprosencephaly Coloboma Sonic hedgehog Gdf6 Retinoic acid Retina a b s t r a c t In this review, we highlight recent literature concerning the signaling mechanisms underlying the development of two neural birth defects, holoprosencephaly and coloboma. Holoprosencephaly, the most common forebrain defect, occurs when the cerebral hemispheres fail to separate and is typically associated with mispatterning of embryonic midline tissue. Coloboma results when the choroid fissure in the eye fails to close. It is clear that Sonic hedgehog (Shh) signaling regulates both forebrain and eye development, with defects in Shh, or components of the Shh signaling cascade leading to the generation of both birth defects. In addition, other intercellular signaling pathways are known factors in the incidence of holoprosencephaly and coloboma. This review will outline recent advances in our understanding of forebrain and eye embryonic pattern formation, with a focus on zebrafish studies of Shh and retinoic acid pathways. Given the clear overlap in the mechanisms that generate both diseases, we propose that holoprosencephaly and coloboma can represent mild and severe aspects of single phenotypic spectrum resulting from aberrant forebrain development. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Introduction to holoprosencephaly and coloboma 2. Part I: Holoprosencephaly As the vertebrate embryo develops, the ventral forebrain and retina are subject to similar extracellular cues that drive their development. A combination of human genetic, animal model, and cell biochemical studies has shed light on the signaling pathways that are essential for patterning this region of the embryo. Within the forebrain, researchers have identified deficiencies in Nodal, Shh, and retinoic acid (RA) pathways, as causes of the ventral forebrain birth defect holoprosencephaly. In the retina, mutations in components of the Shh, RA, and BMP signaling pathways prevent proper closure of the choroid fissure, and result in coloboma. The overlap in genetic causes for these two distinct birth defects implies that they may be more accurately considered as components of a common phenotypic spectrum. This review will define the recent advances in our understanding of Holoprosencephaly (Part I) and Coloboma (Part II), and propose the commonality of their causality. Holoprosencephaly (HPE) is the most common forebrain birth defect in humans [1] and encompasses a broad range of craniofacial and neural defects. Traditionally, HPE has been defined as the failure of the forebrain to divide into two distinct cerebral hemispheres during development. In the most profound cases, the fetus displays cyclopia and the complete absence of lobar divisions in the cerebrum [2–7]. The HPE spectrum also includes ethmocephaly, where the fetus displays a proboscis (an abnormal tubular nose-like structure) and closely-set eyes, and cebocephaly, where the fetus has closely-set eyes and single nostril [8]. More subtle defects have also been classified as falling within the HPE spectrum, including median cleft lip, absence of the olfactory bulb, agenesis of the corpus callosum, and a single, centrally located maxillary incisor [2,8]. While HPE is rare in live births (1/16,000), approximately 1/ 250 conceptuses are thought to be affected by HPE [1,9]. This large difference in frequency likely reflects an incompatibility of severe HPE with life, and a high rate of miscarriage of affected embryos. HPE can be caused by both genetic abnormalities and exposure to teratogens, including alcohol [10,11] and retinoic acid [12,13]. Approximately one-quarter of cases are due to detectable genetic anomalies [5]. Mapping chromosomal deletions in affected patients facilitated the identification of the first genes linked to HPE [14–20], ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Corresponding author. CW405, Biological Sciences Building, Biological Sciences Department, University of Alberta, Edmonton, Alberta, Canada T6G 2E9. Tel.: +1 780 492 4403; fax: +1 780 492 9234. E-mail address: [email protected] (A.J. Waskiewicz). 1 Present address: Laboratoire de Génétique Moléculaire du Développement, U784 INSERM, Ecole Normale Supérieure, 46 rue d'Ulm, 75230 Paris cedex 05, France. 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.09.005 P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 and later work has demonstrated a high frequency of microdeletions and point mutations in these genes [21]. Research in zebrafish over the last decade has yielded many insights into the normal function of HPE genes, and how their disruption can result in human disease. The eight genes linked to HPE have been shown to play important roles in three major developmental signaling pathways: Nodal, Sonic hedgehog, and retinoic acid (RA). The forkhead box transcription factor FOXH1 (also known as FAST1 and schmalspur in zebrafish) [22] and EGF-CFC family protein CRIPTO (also known as TDGF1 in mammals and one-eyed pinhead in zebrafish) [23] are components of the Nodal pathway. The signaling ligand SONIC HEDGEHOG (SHH), transmembrane receptor PATCHED (PTC) and transmembrane sterol-sensing protein DISPATCHED-1 (DISP1) are core members of the Sonic hedgehog signaling pathway, while the transcription factor SINE OCULIS HOMEOBOX-3 (SIX3) directly regulates shh expression [24–27]. The atypical homeodomain transcription factor TG-INTERACTING FACTOR (TGIF) regulates retinoic acid signaling during development [28–30], while ZINC FINGER OF CEREBELLUM (ZIC) transcription factors have recently been proposed as a regulator of Nodal, Sonic hedgehog, and retinoic acid signaling in the forebrain [31]. There is significant crossregulation among these three pathways, which has important implications for early brain development. This cross-regulation is an important consideration in interpreting phenotypes associated with an alteration of gene function within the three pathways, such as in HPE patients (see Fig. 1 for an anatomical overview of forebrain signaling relevant to HPE). 2.1. Specification of the prechordal plate The prechordal plate is the anterior-most axial mesoderm, and a critical ventral patterning center that underlies the developing forebrain. Research in zebrafish has delineated a clear role for Nodal 391 signaling in the formation of the vertebrate prechordal plate. Disruption of this tissue results in the fusion of the eye fields and deletion of ventral forebrain tissue, a phenotype that strongly resembles the most severe cases of HPE [32–36]. The prechordal plate is the anterior-most derivative of Spemann's organizer (the shield in zebrafish), which, during gastrulation, acts as a signaling center to induce dorsal and anterior structures. The shield is specified very early in zebrafish development (the 128 cell-stage) and arises from the cells of the dorsal margin. The transcription factor β-catenin is stabilized in the nucleus of prospective shield cells, and when zygotic transcription initiates, activates transcription of the Nodal ligand and morphogen squint [37]. Within the shield, levels of Nodal signaling are asymmetric. The most vegetal cells are subject to the highest levels of Nodal signaling, and become the prechordal plate, while cells positioned nearer the animal pole, exposed to lower levels of Nodal signaling, become notochord [38]. The prechordal plate, after the cell movements of gastrulation, comes to underlie the ventral forebrain, while the notochord underlies the more posterior neural tube. 2.2. Nodal signaling and HPE Nodal is a member of the TGF-β superfamily of signaling proteins. Members of this protein superfamily exist as dimers, and activate type I and type II transmembrane serine/threonine kinase receptors. Nodals bind the type I receptor protein ALK4 or the complex of ALK7 (type I) and ActRIIB (type II) [39]. The HPE gene CRIPTO/oep physically interacts with the ligand–receptor complex, and is required for Nodal signal transduction, positioning this protein as a Nodal co-receptor [40–43]. Ligand binding results in the phosphorylation of the type I receptors, which in turn phosphorylate the transcription factor Smad2. Phosphorylated Smad2 is imported into the nucleus, where it binds Smad4 and other cofactors, including bonnie and clyde (bon; mixer) and the HPE gene FOXH1 (schmalspur) [44,45]. The Smad complex then activates the transcription of target genes, such as the prechordal-plate-specific transcription factor goosecoid (gsc) [46] and the morphogen shha [47]. Forward genetic screens in zebrafish have identified mutations in the Nodal pathway components linked to HPE: the co-receptor Cripto/ oep, and the Smad2 cofactor FOXH1/sur [48,49]. Loss of function of these genes causes the loss or reduction of the shield and the elimination of the prechordal plate, causing severe forebrain deficits within the HPE spectrum, including cyclopia. 2.3. Cripto/one-eyed pinhead in HPE Fig. 1. Ventral forebrain patterning is regulated by signaling between the pectoral plate and overlying neural tube. Holoprosencephaly (HPE) is caused by a complex set of molecular genetic signaling pathways. Retinoic acid (RA, gradient shown in red) plays a key role in this process and is synthesized in the posterior paraxial mesoderm by Aldh1a2 and degraded anteriorly by Cyp26a1. Tgif, which causes HPE, regulates expression of both cyp26a1 and aldh1a2. Within the prechordal plate (PCP, blue), signaling among Zic, Nodal (Cyclops or Cyc) and Sonic hedgehog (Shh) creates a signaling center underlying the ventral diencephalon. Shh signaling activates both ptc1 and nkx2 genes. oep transcript is maternally deposited in 1-cell zebrafish embryos, and embryos lacking both maternal and zygotic oep (MZoep) have a much more severe phenotype than zygotic mutants. MZoep mutants fail to develop a morphologically distinguishable shield, and lack all shield-specific gene expression. As a consequence, embryos lack most mesodermal cell types, including the prechordal plate and notochord. This in turn causes a failure to specify ventral neurectoderm, and embryos display a cyclopic phenotype [40]. In zygotic oep mutants that retain maternal sources of this protein, shield-specific gene expression is initiated normally, but fails to be maintained. The notochord is specified correctly, and expresses shha, suggesting that low levels of Nodal signaling are sufficient for notochord development. However, zygotic oep mutants lack prechordal plate, and shha is not expressed in forebrain regions, indicating that the specification of prechordal plate is particularly sensitive to a reduction in Nodal signaling [36]. Two human patients with mutations in CRIPTO have been described [23]. One patient displays several HPE-like microforms, including a hypoplastic corpus callosum. This patient has a heterozygous P125L mutation in the CFC domain of the oep orthologue 392 P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 TDGF1, which is required for the protein to bind type I TGF-β receptors. While wild type TDGF1 will rescue the zebrafish oep mutant phenotype, a P125L mutant version does not, suggesting that this protein is nonfunctional, and the malformations in this patient may be due to TDGF1 haploinsufficiency. A second patient displays semi-lobar HPE, and a V159M mutation in CRIPTO, which is in a domain thought to mediate the protein's association with the cell membrane [50]. However, this mutant version of the protein can still rescue an oep mutant, suggesting that the basis of HPE in this case may involve additional genetic or environmental factors. Indeed, within most patient families, the penetrance of HPE is highly variable. It has been widely speculated that interaction between multiple genetic deficiencies or environmental insults are responsible for this variability, and underlie most cases of HPE [4–6]. 2.4. The transcription factor Foxh1 and HPE FOXH1 is a Smad-binding protein which targets activated Smad complexes to specific DNA sequences [51]. It is required for transduction of a subset of Nodal signals, and is partially redundant with the paired-like homeobox transcription factor Bon/Mixer [52]. Zebrafish lacking maternal and zygotic Sur (MZsur) have reduced, but not eliminated, shield-specific gene expression. MZsur embryos also lacking bon have abolished shield-specific gene expression, indicating that these genes are partially redundant. Foxh1 mouse mutants entirely lack specification of the node, prechordal plate, and notochord, and display a severely hypoplastic forebrain [53]. Similarly, MZsur zebrafish mutants also show the elimination of the prechordal plate, although notochord-specific gene expression is still present. Interestingly, notochord-specific genes are activated even in the absence of both bon and sur, suggesting that these two are not the only Smad cofactors that can transduce Nodal signal to specify notochord in zebrafish. Besides its role as a Nodal signal transducer, Sur is also required for the maintenance of cyclops and squint expression during gastrulation [44], however, expression of neither cyclops nor squint can rescue prechordal plate specification, cyclopia, or other phenotypes in MZsur embryos. Therefore, it is likely that these defects mostly result from an inability to transduce Nodal signals within the axial mesoderm. A number of mutations in FOXH1 have been identified in HPE patients [22], including those that disrupt the forkhead DNA binding domain and Smad-interaction domain. The activity of these mutant versions has been assessed by measuring the activity of RNA to rescue zebrafish sur mutants. These studies demonstrate that mutations in the Smad-interacting domain eliminate protein activity, while variants in the forkhead domain reduce protein activity by approximately 20% [22]. Overall, in live births, the occurrence of Nodal pathway mutations is rare [22]. Work in animal models suggests that this is likely due to the essential nature of this signal pathway during the very early stages of development. Indeed, in patients suffering from recurrent miscarriage, a very high proportion of embryos lost in the first trimester show severe neural and midline defects [54]. Embryos with severe loss of Nodal function are likely not viable, and it is plausible that Nodal pathway mutations contribute to a proportion of early pregnancy losses. 2.5. Sonic hedgehog signaling and HPE Defects in the Sonic hedgehog pathway are the most frequent cause of HPE [18], and this pathway has long been known to be essential for forebrain patterning. An essential role for Nodal signaling is the initiation of shh transcription in the prechordal plate and notochord [38]. Shh secreted from the axial mesoderm initiates shh transcription in the floor plate, the most ventral and medial cells of the neural tube. A graded hedgehog signal emanating from both the axial mesoderm and floor plate is required for diencephalon specification, ventral neuronal identities, and the correct separation of the eye field into two distinct domains [55]. The mechanism of Shh signal transduction has been the subject of intense research. The protein undergoes extensive post-translational modifications, including autocatalytic cleavage and cholesterylation of the N-terminus, which becomes the mature ligand [56]. Subsequently, a palmitate moiety is added to the N-terminus of the ligand by the enzyme known as Hedgehog acyltransferase or Skinny hedgehog [57]. The N-terminal signaling ligand processed with both lipid modifications is denoted as ShhNp. Cleavage and lipid modification is required for secretion outside the cell. A series of proteins, including Dispatched and You/Scube2 are required for its export. ShhNp then binds to transmembrane proteins (Interference hedgehog; IHOG and Brother of IHOG; BOI) that facilitate its interaction with the Patched receptor [55]. In the absence of ligand, Ptc inhibits the function of Smoothened (Smo). Upon ligand binding, this inhibition is relieved; Smo becomes activated, and induces a signaling cascade that leads to the regulation of the localization or cleavage of Gli transcription factors. In Drosophila, a single Gli protein, named Cubitus interruptus (Ci), is present. Upon ligand binding, the cleavage of Ci into a repressor isoform is inhibited, and the activator isoform transcriptionally upregulates target genes [58]. Sixty-four unique mutations in the SHH gene have been linked to holoprosencephaly [59]. These include frameshift and nonsense mutations, missense mutations in functional domains or uncharacterized but conserved domains, or changes in residues important for processing, such as the post-translational addition of palmitate and cholesterol moieties. Two orthologues of SHH exist in zebrafish, shha and shhb (formerly known as tiggy-winkle hedgehog). This complicates analysis of mutants for these two genes, as each can compensate for the other's function. Mutants lacking shha have a mild phenotype, including abnormal fin and somite patterning, and axonal pathfinding defects, while knockdown of shhb alone causes no detectable phenotype [60–62]. Because of redundancy, shh loss of function has mostly been accomplished in zebrafish by mutations in or pharmacological inhibition of Smoothened (Smo) [63]. The compound cyclopamine directly binds and inhibits Smo signal transduction [64]. Unlike the Shh ligand itself, smo does not appear to be duplicated in the zebrafish genome, and smo mutants show a severe phenotype, although not as severe as cyclopamine-treated embryos. Smo is present maternally, and maternal transcript is likely to allow for some early Shh signal transduction, which cyclopamine would block. smo mutants have a suite of defects, including synopthalmia and improper motor neuron specification in the spinal cord. Moreover, mutants show a dorsalized forebrain and a complete lack of hypothalamic tissue. smo mutants also have a lack of anterior pituitary cell fates, and instead form an ectopic lens at this site. Interestingly, an optic stalk marker, pax2a, is ectopically expressed in the forebrain, suggesting that hedgehog signals play a key role in distinguishing forebrain from eye identities [63]. The HPE gene PATCHED is the receptor for Shh, and is in fact a negative regulator of the pathway. Unliganded Ptc represses Smo, and ligand binding prevents Ptc repression activity. Therefore, loss of Ptc function results in constitutive activation of the Shh pathway [65–67]. Zebrafish have two ptc homologues, and mutants for both ptc1 and ptc2 have a strong upregulation of hedgehog signaling, concomitant with a ventralization of the neural tube and mispatterning of the somites [68]. Conversely, gain of Ptc function results in constitutive repression of the pathway, and also results in phenotypes within the spectrum of HPE, including midline defects, such as malformation of the corpus callosum [66,69]. Several PTC mutations have been linked to HPE. Two of these are in the extracellular domain in a position thought to interact with the SHH ligand. Two others are located within intracellular domains of the protein, and it is hypothesized that these domains might be important for PTC's interaction with SMO [70]. It has been proposed that these P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 mutations represent a gain of PTC inhibitory function of the SHH pathway, as HPE phenotypes are well-known to be linked to SHH loss of function. However, it is possible that a gain of SHH function could cause HPE-like phenotypes as well, since in zebrafish, a loss of Ptc1 and consequential increase in signaling results in narrow-set eyes and moderate forebrain patterning defects [68]. The function of the HPE gene DISPATCHED has also been modeled in zebrafish, where there is a single described orthologue [71]. Although disp1 expression is broad, it is enriched in regions of high Shh signaling: the ventral forebrain, notochord, and floor plate. In Drosophila, disp has been shown to be required for export of processed Shh ligand [72]. Suggesting that this function is conserved, disp lossof-function in mice and zebrafish causes phenotypes typical of Shh deficiency [73–76]. Zebrafish embryos with mutations in disp1 (chameleon) have phenotypes typical of a reduction in Shh signaling: strongly reduced expression of shh target genes, and a dorsalized neural tube. Further supporting the model that disp regulates Shh export, overexpressing shh RNA in disp1 zebrafish mutants causes an attenuated transcriptional response compared to shh overexpression in wild type embryos [76]. 2.6. The transcription factor Six3 and HPE Zebrafish models of the function of the HPE gene SIX3 are complicated by the presence of three SIX3-related genes, six3a, six3b, and six7, which are all expressed in the prechordal plate during gastrulation, and then in the anterior neurectoderm during somitogenesis [77]. Interestingly, however, six3b/six7 loss of function indicates that Six3 proteins are required in the neurectoderm to repress Nodal target gene expression, which contributes to the proper establishment of left/right asymmetry [77]. Complete loss of six3 gene function may reveal additional roles for six3 regulation of Nodal signaling. Zebrafish rescue studies indicate that the vast majority of HPEassociated SIX3 mutations act as hypomorphs [78]. Work in mice suggests that defects in a direct interaction between SIX3 and the Shh signaling pathway underlie the development of HPE [24]. Geng and colleagues generated mice with a HPE mutant version of Six3 knocked-in to the endogenous Six3 locus. While heterozygous mutant mice display HPE-like phenotypes at a low frequency, removing one copy of Shh results in severe cyclopia and loss of ventral diencephalic gene expression. Further, mice heterozygous for the mutant Six3 and with one copy of Shh deleted have a loss of both Shh and Six3 expression in the ventral diencephalon, indicating that these two genes form a critical positive feedback loop. The regulation of Shh by Six3 is direct, and the Six3-binding site in the Shh forebrain enhancer is mutated in a patient with semi-lobar HPE [79], indicating the importance of the genetic interaction between these two proteins for forebrain development. 2.7. Retinoic acid signaling and HPE Retinoic acid has a critical role in early neural patterning. This small, diffusible molecule is synthesized from vitamin A (retinol) precursors through a series of enzymatic reactions. The retinol dehydrogenase Rdh10 catalyzes the conversion of retinol to retinaldehyde during development [80], and aldehyde dehydrogenases (raldh or aldh1a gene family) convert retinal to retinoic acid [81]. There are two characterized aldh1a genes in zebrafish: aldh1a2, which is expressed at high levels in paraxial mesoderm and in the dorsal retina [82,83], and aldh1a3, which is expressed in the ventral retina flanking the optic stalk [84]. Recently, an additional enzyme, cyp1b1, has been demonstrated to synthesize RA in multiple tissues during chick development [85]. This gene is expressed in a much more restricted domain in zebrafish than chick, and is present in the anterior telencephalon and eye, beginning mid-somitogenesis (Gongal et al., submitted). Its 393 contribution to zebrafish neural or eye patterning has yet to be established. However, mutations in mouse and human Cyp1b1 are linked to the development of juvenile-onset glaucoma, suggesting an important role in eye development [86–88]. The best characterized role of RA during development is in the hindbrain, where it acts as a potent posteriorizing signal on neural tissue. High levels of RA signaling are required for specification of caudal hindbrain fates [82,83,89–91]. During gastrulation, the forebrain is insulated from the posteriorizing influence of RA through the activity of cyp26 (cytochrome P450, subfamily XXVI) genes, which hydroxylate RA, a modification thought to target the molecule for degradation [92]. cyp26a1 is the earliest expressed cyp26, and is expressed in the anterior neurectoderm, initiating during gastrulation [93]. Slightly later, cyp26b1 and cyp26c1 transcription is initiated in a subset of telencephalic cells (cyp26c1) and the diencephalon (cyp26b1 and cyp26c1) [94,95], where they limit RA activity. Both cyp26b1 and cyp26c1 are also expressed in dynamic, rhombomere-specific patterns in the hindbrain, where they are essential for proper anterior– posterior patterning [94–97]. Together, RA synthesis and degradation genes set up a highly dynamic pattern of RA levels in the developing CNS. Abnormal RA levels have severe phenotypic consequences, particularly for the forebrain. Exposure to exogenous RA during development causes severe craniofacial defects, eye patterning anomalies, and ectopic forebrain expression of posterior CNS genes [11,98,99]. RA levels become abnormally high when Cyp26 activity is reduced, which has been modeled in mice, chick, and zebrafish. Overall, embryos with reduced Cyp26 function display defects in patterning of the diencephalon, eye, hindbrain and neural tube closure defects [30,96,100–102]. Further, without Cyp26 activity, embryos are highly sensitive to environmental changes in retinoids, and the entire neural tube takes on posterior hindbrain identity upon exposure to doses of retinal that cause no defects in wild type embryos [96]. High levels of retinoic acid have recently been linked to DiGeorge syndrome (also known as velo-cardio-facial syndrome), a suite of multiple congenital defects, including craniofacial abnormalities such as cleft palate [103,104]. A critical region mapped in patients with chromosomal deletions encompasses the gene tbx1, which regulates expression of all three cyp26 genes during somitogenesis [102]. Taken together with the large body of evidence that increased developmental RA levels can cause forebrain patterning defects, it is plausible that changes in RA signaling underlie some of the phenotypes observed in DiGeorge syndrome. RA deficiency also causes severe developmental defects. Loss of RA has been modeled in vitamin A-deficient quail and swine [105–107], mouse and zebrafish mutants for RA-synthesizing genes [82,83,90,108], and chick and zebrafish by application of pharmacological inhibitors [89]. Although defects observed vary slightly depending on the organism, common phenotypes include the formation of a single telencephalic vesicle, mispatterning of the diencephalon, failure to specify the caudal hindbrain, and eye abnormalities. Thus, both too much and too little RA cause defects in similar tissues, and many of these phenotypes are reminiscent of HPE. Interestingly, similar defects are observed in a gain or loss of RA, such as the mispatterning of the diencephalon and failure of the neural tube to close. 2.8. The transcription factor TGIF and HPE Until recently, little was known about the role of the HPE gene TGIF (5′-TG-3′ interacting factor) in forebrain development. Work in cell culture systems demonstrates that TGIF can transcriptionally regulate both retinoid and TGF-β signaling. Tgif can bind Smad2 and Smad3, and represses the transcription of a TGF-β-dependent reporter [109], and mutant versions of TGIF that occur in human patients do not show 394 P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 this same repression activity [110]. As alterations of Nodal signaling are well-known to cause forebrain defects in animal models, and Nodal is a TGF-β pathway, it was widely assumed that misregulation of Nodal function explained the incidence of HPE in human patients deficient in TGIF [5,7,110–113]. However, there has been a notable absence of evidence from animal models for this hypothesis. In fact, TGIF was originally identified because of its involvement with retinoid signaling. This atypical homeodomain transcription factor was discovered in a search for proteins that bind to a retinoid X response element (RXRE) [29]. Bartholin and colleagues provide further evidence that TGIF can regulate retinoid signaling [28]. They convincingly show that TGIF binds RXRs (most robustly RXRα and RXRγ, and RXRβ to a lesser extent), and that TGIF can form a complex both with RXR and its nuclear receptor binding partners, including RARs and PPARs (peroxisome proliferator activated receptor). Using a reporter assay, Tgif can repress RA-dependent transcription under the control of a DR5 element. However, this element is activated by RXR expression, regardless of the presence or absence of ligand. Further, it isn't clear whether this repression activity is specific for RXR or generally acts on other nuclear receptors. Therefore, the in vivo relevance of Tgif's function in this assay remains to be established. Four independent groups have generated Tgif-null mice [28,114– 116]. None of these mutants has neural patterning phenotypes. Mice have multiple copies of Tgif-like genes, including Tgif2, Tgif-like-onthe-X and Tgif-like-on-the-Y, whose functions are poorly understood, and which may compensate for the loss of Tgif activity in these mutants. However, depending on genetic background, heterozygous and homozygous mutants do have an increased rate of exencephaly when treated with RA [28,116], suggesting that Tgif plays a role in the regulation of RA signaling. A fifth group interested in Tgif function noted that some patients had a nonsense mutation that left the first repressor domain and homeodomain intact, but eliminated the second and third repressor domains. Kuang and colleagues generated a mouse with an equivalent mutation by deleting Tgif's third exon. Interestingly, mice with this mutation do show neural patterning defects, although these are strain-specific [117]. TgifΔexon3 mice show severe hypoplasia of the forebrain and exencephaly. Shh protein levels are reduced, and the ventral forebrain marker Nkx2.2 has reduced expression. Because of these phenotypes, Kuang and colleagues propose that this version of Tgif may function as a dominant negative. However, defects in this mutant are still highly dependent on genetic background, which suggests that there are unknown and essential genetic modifiers that exist in this strain. In contrast to mice, zebrafish have only two tgif family genes, tgif and tgif2, and therefore genetic redundancy may be less of an issue for functional analyses. Studies in zebrafish have revealed that Tgif is an essential regulator of retinoic acid metabolism gene expression [30]. Tgif is required for the initiation of both cyp26a1 and aldh1a2, indicating that this gene regulates both RA degradation and synthesis. tgif morphants have reduced diencephalic gene expression, a phenotype that resembles that of cyp26a1 mutants, suggesting that excess RA in the forebrain is responsible for this mispatterning. These results suggest a molecular basis for Tgif-dependent holoprosencephaly; embryos lacking TGIF function may be highly sensitive to teratogenesis from environmental fluctuations in vitamin A [30]. 2.9. Genetic interactions between Sonic hedgehog and retinoic acid signaling in the forebrain While deleterious for the forebrain during gastrulation, RA is in fact required for proper forebrain development at later stages. One of its important roles is facilitating Sonic hedgehog signaling. The genetic interactions between Shh and RA signaling are complex, and are quite different between species, as well as between different tissues and developmental stages (see below). RA regulates Shh signaling at multiple points in the pathway, including transcription of the ligand, response to the signal, and post-translational processing. The mechanisms of these interactions are unclear, and identifying the precise molecular bases that underlie RA-Shh interactions are key topics for future work. Interpreting the effects of changes in shh expression can be challenging, as transcription of this gene is regulated by an autoregulatory loop. When Shh signaling is reduced, shh transcription is upregulated [118], so both upregulation and downregulation of the pathway may be correlated with an increase in shh transcription. Nevertheless, shh expression is critical for activity of the pathway, and RA regulates signaling at this level. Aldh1a2 mouse mutants have normal Shh transcription (and protein levels, detected by immunohistochemistry) in the axial mesoderm during early stages of forebrain patterning. Slightly later in mouse development (by the 16 somite stage), however, RA does regulate the transcription of Shh itself. This regulation is tissuespecific, however, as in a Aldh1a2 mutant, Shh is decreased in some tissues (ventral diencephalon), and increased in others (infundibulum) [108]. The functional consequences of this loss of transcript are unclear, but occur too late to explain early loss of Shh target gene expression in aldh1a2 mutants. In contrast, transcriptional regulation of shh by RA seems to be more important in bird development. In vitamin A-deficient quail, shh expression is reduced in prechordal plate at early stages [106]. Further, locally applying RAR and RXR inhibitors in the chick forebrain abolishes shh expression. Treated embryos display cyclopia and loss of forebrain tissues [119]. This forebrain phenotype is more severe than an Aldh1a2 mouse mutant, perhaps because receptor inhibition causes a more profound loss retinoid activity. Aldh1a2 and Aldh1a3 are partially redundant in the mouse forebrain patterning, but double mutants cannot be analyzed (without maternal supplementation with RA), due to very early lethality caused by cardiac defects. A second method of RA regulation of Shh signaling has recently been revealed in mouse. Aldh1a2 mutant mice have a reduction in Shh target gene expression, such as Gli3 and Olig2, while no change in Shh transcription or protein levels is detectable [108]. Importantly, Ribes and colleagues demonstrated that RA deficiency causes a failure to appropriately respond to the Shh signal [120]. Although the mechanism of this interaction is not yet understood, treating Aldh1a2 mutants in vitro with Shh ligand does not result in high levels of Shh target gene expression, as does treating wild type embryos. These results could reflect a role for RA in regulating the activity of Shh protein activity, localization, processing, degradation, or RA signaling could be required for the transcription or function of a Shh pathway component. Future experiments will reveal the myriad of mechanisms used in regulating Shh activity, such as modulation of palmitoylation [121,122]. 2.10. Zic genes and the coordinated regulation of Shh and RA signaling While cross-talk between the Shh and RA pathways is likely essential to coordinate their activities during neural patterning, these two pathways are also connected by common regulators. A transcription factor that regulates both the RA and Shh pathways has been identified: zic1. The precise effect of a single factor transcriptionally regulating components of multiple pathways is not well understood. However, it is plausible that this method of regulation functions to maintain a precise balance between the activities of the two pathways. zic1 (whose related gene zic2 has been linked to HPE) has been linked to disruptions to both the retinoic acid and Shh signaling pathways [31]. zic1 morphants show a loss of Shh target gene expression (ptc1), concomitant with a mild reduction in shha expression and a strong reduction in shhb expression. zic1 loss-of-function results in higher levels of RA signaling in the eye and forebrain mid-somitogenesis. zic1 morphants have an expansion of optic stalk in the eye and P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 forebrain, a tissue that strongly expresses the RA-synthesizing genes aldh1a3 and cyp1b1, which might cause the high RA levels observed in the eye and forebrain. Overexpressing cyp26a1 makes some zic1 morphant phenotypes more severe, suggesting that RA deficiency, rather than excess, contributes to these phenotypes. Indeed, RA levels in the somites are reduced, as assayed by RARE-yfp reporter expression. Nevertheless, although the detailed mechanisms underlying Zic1 function are not yet clear, its overall function is clearly to regulate both RA and Shh signaling, which may help facilitate the coordinated actions of these two pathways during neural development. Zebrafish zic2 has also been shown to play a key role in regulating forebrain development. Morphant embryos that lack zic2a display a pronounced reduction in arx and dlx2 gene expression, indicating a defect in specification of the prethalamic region of the forebrain [123]. Though the changes in ptc1 gene expression are mild, it is possible that other zic genes are compensating for a loss of zic2a [124]. Recent research on zic2a has shown that it plays an essential role in regulating proper formation of the eye. Indeed many of the genes discussed in this section function at the same developmental stage to regulate formation of the eye. Given the striking similarities between forebrain and eye patterning mechanisms, the second component of this review will deal with eye formation birth defects, most notably colobomata, and the known signaling pathways that regulate ocular morphogenesis. 3. Part II: Coloboma During early development, the vertebrate eye contains a transient ventral fissure that is essential for the entry of mesenchyme cells into the retina, where they will give rise to blood vesicles that will nourish the eye throughout the life of the organism [125]. Failure of the optic fissure to close results in coloboma, an important cause of visual impairment worldwide. Depending on the size and location, this congenital malformation can lead to blind spots, lazy eye, and reduced visual acuity in patients [126]. Coloboma can affect numerous tissues in the posterior eye including the choroid, and optic nerve, as well as tissues of the anterior eye such as the iris, cornea, and lens [127]. Ocular colobomata have been observed as an independent defect, or as a component of syndromes that often include other ocular abnormalities as well as craniofacial and neurological phenotypes [128]. These include microphthalmia [127,129,130], cataracts [131], and CHARGE syndrome (coloboma, heart defects, choanal atresia, retarded growth, and genital and ear anomalies) [132]. Colobomata can be sporadic, or heritable with autosomal dominant, recessive, and X-linked inheritance cases being reported. The vertebrate eye is patterned along proximal–distal axis early in development. The distal eye cells give rise to neural retina and retinal pigmented epithelium, while more proximal cells give rise to the optic stalk and choroid. Although transient, the choroid fissure acts as an entryway into the eye, allowing vascularization to occur. Before the onset of cellular differentiation in the retina, the fissure begins to close using a contact-dependent dissolution of the basal lamina at the contacting neuroepithelial on each side of the optic fissure [133]. Thus, genetic abnormalities or environmental toxins that disrupt this process, likely present as coloboma. A number of genetic lesions and chromosomal rearrangements have been identified that lead to coloboma in human patients, many of which involve the loss of Sonic hedgehog signaling or one of its downstream target genes [134,135]. As aberrant Shh signaling is one of the leading causes of holoprosencephaly, it is important to note that many individuals suffering from mild forms of HPE also display coloboma. In addition, defects in the RA [136,137], and Bmp [127] signaling pathways have been implicated in the generation of coloboma or syndromes involving colobomataneous phenotypes. 395 3.1. Sonic hedgehog signaling is required for closure of the optic fissure To date, only one family has been identified with coloboma due to mutation in the sonic hedgehog (shh) gene itself [135]. It is likely that null mutations, or even strong hypomorphic mutations in the Shh gene are incompatible with life as Hedgehog signaling is required forebrain patterning and development. Thus, most knowledge concerning Shh and its role in ocular coloboma comes from experiments in animal models. Both the zebrafish and mouse have proven to be useful models for the study of the eye, and have provided valuable insight as to the role of Shh and its downstream target genes in the development of coloboma. In zebrafish, like most other vertebrates, there are two tissues that act as sources of Sonic hedgehog ligand that are relevant for the development of the eye: first, the floorplate in the ventral midline of the diencephalon, and second, the neural retina itself. It is the former that is responsible for eye patterning and closure of the optic fissure [138], while Shh signaling derived from within the retina is required later in development for the timely cell cycle exit and differentiation of neural lineages [139]. 3.2. Sonic hedgehog regulates genes required for eye patterning The zebrafish eyecup is partitioned early in development, establishing the retina, optic stalk, and retinal pigmented epithelium into molecularly distinct domains. Midline diencephalic Sonic hedgehog is required for this partition by maintaining the expression boundary of spatially restricted transcription factors [138]. Loss of shh in zebrafish leads to severe coloboma, and the loss of ventral-anterior homeobox 1 and 2 (vax1 and vax2), which delineate the optic stalk and ventral neural retina regions [140]. Inactivation of both vax genes in zebrafish results in severe coloboma [140], supporting the model that shh acts genetically upstream of vax genes to regulate closure of the optic fissure (Fig. 2). In addition to regulating vax transcription factors, Shh is required for the proper expression of members of the paired box family of transcription factors, with pax2 and pax6 being important for eye development and closure of the optic fissure. Shh is required for expression of pax2, while negatively regulating the levels of pax6 [138,141]. Mutations in both genes have been mapped in patients with coloboma or colobomataneous syndromes; mutations in pax2 are linked to renal-coloboma syndromes [142,143] while mutations in pax6 tend to be found in patients with coloboma of the optic nerve (along with other retinal phenotypes including aniridia). pax2 expression is restricted to proximal eye structures such as the optic stalk and fissure, while pax6 is restricted to more distal structures. The expression of pax2 in cells surrounding the choroid fissure requires Shh, and overexpression of shh leads to pax2 expression in more distal structures such as the neural retina [138]. Conversely, loss of Shh signaling results in reduced expression of pax6 neural retina. Thus, it appears that control of pax gene expression via shh is required for the proper partition of early eye to define and maintain the partition of optic stalk and neural retina. It is clear from studies involving both Pax and Vax proteins, that early embryonic patterning of the eye into distinct optic stalk and neural retinal lineages is essential for subsequent fusion of the optic fissure. Recent evidence from zebrafish model studies suggests that the regulation of pax genes via Sonic hedgehog signaling is indirect, and involves the regulation of other factors that modulate pax gene expression. One such factor, Zinc finger protein of the cerebellum 2a (Zic2a), has been shown to negatively regulate the transcriptional output of Sonic hedgehog signaling, while itself being transcriptionally regulated by Hedgehog signaling cascades [124]. Loss of zic2a results in severe coloboma [124,144,145] likely due to mispatterning of the optic stalk and retina, as optic stalk markers are ectopically expressed in the retina in Zic2a-depleted embryos [124]. 396 P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 tion of the retinoblastoma gene (Rb), leading to abnormal cell cycle progression and increased proliferation [147]. Work in mice demonstrates that accurate regulation of the cell cycle acts upstream of Sonic Hedgehog and causes abnormalities in the eye. Loss of function studies of involving the cell cycle regulator E2F4 result in coloboma, and aberrant expression of Sonic Hedgehog ligand at the ventral midline [148]. E2F4 is a transcription factor downstream of the Rb gene, further linking Rb cell cycle control to the development of coloboma. In addition, Shh effector genes are abnormally expressed upon loss of E2F4, with Pax2 expression extending ectopically to the distal optic cup, and reduced Pax6 expression in the ventral optic cup [148]. 3.3. Retinoic acid plays a key role in closure of the optic fissure Fig. 2. Sonic hedgehog signaling is required for eye patterning the eye along the mediolateral axis. A complex transcriptional interaction between Sonic hedgehog (Shh) and Zic pathways specifies the identity of the midline of the ventral diencephalon. The Shh diffusible ligand is required for specification of ventral eye structures, including the optic stalk and ventral neural retina. Shh is required for expression of vax2 and pax2 transcription factors, both of which are expressed in the optic stalk and ventral retina. This signaling cascade is opposed in the dorsal retina through the action of Bmp ligands, such as Gdf6. Bmp signaling cascades are required for the expression of dorsal retinal specific transcription factors such as Tbx5 and Tbx2. Loss of Shh signaling, or any of its downstream signaling components can lead to coloboma. Loss of Bmp signaling in the dorsal retina can also lead to coloboma, as well as ectopic fissure formation in the dorsal retina. Although numerous candidate genes have been identified downstream of Shh for the development of coloboma, there is little data concerning the mechanism by which fusion of the optic fissure is prevented in shh mutants. In wild type animals, fusion is facilitated by contact between the epithelial tissue on either side of the optic fissure leading to a degradation of the basal lamina. Inhibition of Shh signaling and the consequential mispatterning along the proximal– distal axis of the eyecup have been shown in some cases to physically prevent contact between epithelium on either side of the optic fissure. For example, in the zebrafish blowout mutant where there is a mutation in ptc1, (a gene that functions as a Shh receptor and negative regulator of Shh signaling; see Section I of this review), ectopic expression of optic stalk markers such as pax2, is observed, and the stalk grows ectopically into the retina, physically preventing contact between the epithelial tissue on both sides of the optic fissure [146]. Conversely, when pax2 is lost in zebrafish, contact between the epithelial tissues on either side of the optic fissure still occurs, yet there is no dissolution of the basal lamina and the fissure persists. Thus, pax2 must be required for fissure closure subsequent to contact between the two sides of the fissure. It has recently been demonstrated that cell cycle regulation may also play a key role in regulation of optic fissure closure, and is linked to Shh signaling. For example, mouse humpty dumpty mutants display severe coloboma and cell cycle defects. This transgenic mouse contains a mutation in Phactr4, which results in hyperphosphoryla- Retinoic acid has long been implicated in the development of the eye in humans and animal models. Like Sonic hedgehog, mutations in components of the core RA pathway are rarely found in human patients with coloboma due to the wide array of functions retinoic acid performs in the developing embryo. However mutations in Retinol Binding Protein (rbp) have been described in a sibling-pair that presented with iris coloboma [149]. In both siblings, no serum Rbp was detected, and patients had approximately 17% of normal retinol levels. In addition to iris coloboma, these patients also suffered from night blindness but were otherwise healthy. Furthermore, there is evidence that suggests that in some populations, vitamin A-deficient (VAD) diets are correlated with the incidence of coloboma [136], reiterating the link between aberrant retinoid levels to the development of coloboma. Animal models have yielded major insights into the role of retinoic acid during eye development. For example, mice that receive teratogenic doses of retinoic acid while pregnant give birth to pups with eye defects, including coloboma [150,151]. In zebrafish, local delivery of RA to the eye via implanted RA soaked beads not only results in coloboma, but can also induce ectopic fissure formation anywhere in the retina [152]. Furthermore, failure to close the optic fissure in RA treated mice correlates with changes in gene expression along the proximal–distal axis of the developing eye [151] in a manner comparable to that seen with a loss of Sonic hedgehog signaling. Epistasis analysis indicates that sonic hedgehog itself is downregulated by teratogenic doses of RA [153], although this relationship was only tested in tissues other than the floorplate (the critical source of Shh ligand for optic fissure closure). Nevertheless, this demonstrates that RA can act upstream of Shh in a living organism. Although teratogenic doses of RA clearly regulate Shh signaling and cause coloboma, little work has been done involving manipulation of RA at physiologically relevant levels. In most studies, embryos are incubated in a concentration of RA that far exceeds what an embryo would naturally be exposed to during development. In the developing embryo, RA levels are higher in the ventral retina than in the dorsal retina, thus creating a gradient of retinoic acid signaling across the dorsal–ventral axis of the eye [154,155]. This gradient cannot be mimicked by simply exposing the whole embryo to RA. Thus, loss of function studies are required to determine if endogenous RA is required for closure of the optic fissure. These studies have been fewer in number, and the role of endogenous RA in fissure closure appears to differ among different model systems. It has been noted, for example, that treatment of developing mice with citral, a competitive inhibitor of aldehyde dehydrogenases [156–158] caused a delay in (although did not prevent) the closure of the optic fissure [151]. Vitamin A depletion in mice also leads to coloboma that occurs with a reduction in ventral retinal cell number and a loss of cell adhesion molecules surrounding the optic fissure [159]. This phenotype may be due to abnormal morphogenesis of the early eyecup, as opposed to later patterning events, as when aldehyde dehydrogenase genes are knocked out in mice, the dorsal eyecup P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 invaginates, but the ventral eyecup completely fails to form [160]. Similarly, in zebrafish, treatment of embryos with citral results in the failure of the ventral eye to form, while the dorsal half remains relatively normal [154]. Knockout of the dorsal-specific aldh1a2 does not result in coloboma, although it is likely that compensation by the ventral-specific aldh1a3, for which there is currently no mutant zebrafish strain available. Thus it appears in both mice and zebrafish, endogenous RA may affect closure of the optic fissure not because of improper patterning of ocular tissues, but due to its requirement for earlier morphological events leading to morphologically normal eyes. 3.4. Bmp signaling and coloboma Mutations or chromosomal aberrations involving members of the bone morphogenetic protein (Bmp) family have also been described in human patients with coloboma. Asai-Coakwell and colleagues demonstrated that a deletion encompassing GROWTH AND DIFFERENTIATION FACTOR 6 (GDF6/BMP13) is likely responsible for bilateral retinal coloboma and unilateral iris coloboma [127]. Further studies revealed that mutations in the prodomain of GDF6 account for 1.6% of retinal colobomata in the human population [161]. This mutation results in reduced Gdf6 protein secretion from the cell, which is mechanistically consistent with Gdf6's non-cell autonomous effects [162,163]. Mutations or chromosomal aberrations involving the closely-related bmp4 can also cause coloboma in human patients [164]. Animal models have confirmed the association of gdf6 and bmp4 with coloboma, and have provided insights as to the mechanism by which each gene can affect closure of the optic fissure. Both genes are expressed in the retina during early stages of development, and knockdown of gdf6a (one of two gdf6 paralogues in the teleost genome) in zebrafish results in retinal coloboma [127,162]. Similarly to exogenous RA application, loss of Gdf6a signaling in zebrafish leads to ectopic fissure formation [162,163]. Studies of axial patterning molecules expressed in the retina indicate that gdf6a is responsible for limiting ventral eye marker expression, as loss of this protein results in expansion of vax2 expression to include both the ventral and dorsal retina. This may explain the ectopic fissure in the dorsal eye, which is respecified to a ventral fate in the absence of Gdf6a. In addition, Gdf6a is required for proper establishment of RA signaling in the eye. For example, the dorsal-specific aldh1a2 is not expressed in embryos lacking Gdf6a, while the ventral-specific aldh1a3 is upregulated [162]. While details regarding the mechanism by which bmp4 can cause coloboma are not as well understood, it has been noted that bmp4 expression overlaps with both sonic hedgehog and patched1 expression. Further suggesting a genetic interaction between BMPs and Sonic hedgehog signaling is that an increased phenotypic severity is observed in human patients with low-penetrance mutations of both bmp4 and shh than with either single mutation [164]. Supporting this model, studies in mice have implicated Bmp4 as essential for Hedgehog induced coloboma. In a conditional knockout of the Smoothened gene, which is required to transduce the Hedgehog signal, Bmp4 expression is strongly expanded throughout the ventral retina [165]. As Bmp4 expression is normally confined to the dorsal retina, the authors speculate that coloboma in these embryos may be caused by ectopic BMP signaling. In addition, the expression of ventral retinal and optic stalk markers, including Vax2 and Pax2, show reduced expression in Smoothened-null embryos, however only at developmental time points after ectopic expression of Bmp4 is observed. Thus, loss of ventral retinal and optic stalk gene expression, and the resulting coloboma in embryos suggest that the lack of Hedgehog signaling results in the inability to limit Bmp4 signaling. Furthermore, overexpression of Bmp4 in the developing optic vesicles can inhibit Hedgehog expression itself [166], supporting a model whereby localized expression of Sonic hedgehog and Bmp ligands 397 constitute antagonizing signaling centers required for patterning the eye along the dorsal–ventral axis and closure of the optic fissure. In many model systems, Bmp4 and Gdf6 have been proposed to function redundantly. As loss of gdf6a results in coloboma and has been shown to regulate retinal expression of bmp4 [162], it is likely that Gdf6a also plays a role in the formation of coloboma in embryos lacking Smoothened function. Surprisingly, loss of Smoothened does not result in any change of expression retinoic acid synthesizing genes such as Aldh1a2 or Aldh1a3 [165]. Gdf6 has been shown to regulate the expression of both of these genes, demonstrating that although Bmp4 and Gdf6a act redundantly for some developmental processes, at least some independent functions for each molecule are required for closing the optic fissure. 3.5. Coloboma may be linked to overall growth of the eye Mutations in human BMP4 and GDF6 result in both coloboma and microphthalmia, indicating that closure of the optic fissure may be genetically linked with the overall growth of the eye in utero. A number of other genes have been reported to cause both microphthalmia and coloboma, including transcription factors such as orthodenticle homeobox 2 (otx2) and Sex determining region Y-box 2 (sox2). Whole gene deletions in Otx2 tend to result in severe microphthalmia, while milder mutations tend to result in either unilateral or bilateral coloboma [167–169], suggesting that the level of gene activity influences both growth of the eye and closure of the optic fissure. Although required for differentiation of the retinal pigmented epithelium [170,171], otx2 expression is also observed in the early proliferating cells of the eyecup [172], and could thus affect growth of the eye before it is required for specification of RPE. However to date, there is no data that sheds light on the mechanism by which otx2 influences both growth of the eye and closure of the optic fissure. Mutations in the SOX2 gene have also been identified in human patients [173]. Similar to Otx2, the severity of loss of function in Sox2 correlates with the severity of disease, whereby mutations causing complete loss of function result in severe microphthalmia, while milder variants tend to manifest as coloboma [167,174]. Although it is not known whether Sox2 is directly involved in closure of the optic fissure, the product of this gene is required for the proliferation of early retinal progenitor cells [174], reiterating the potential genetic link between eye growth and closure of the optic fissure. 4. Conclusions The genetic pathways that influence the formation of HPE and coloboma, as outlined in this review, have considerable overlap. Holoprosencephaly and colobomata represent birth defects that can arise from aberrant Sonic hedgehog signaling. Shh is required in the forebrain for the formation of ventral neural structures and division of the early eye field into two distinct hemispheres. Interference with this signaling pathway, either through genetic mutation or via environmental factors can thus cause holoprosencephaly. Forebrain Shh signaling is also needed for early eye patterning, a process that is required for closure of the optic fissure. Given the overlap in mutational etiology, holoprosencephaly and coloboma are likely to represent severe and mild ends of the same phenotypic spectrum. Retinoic Acid signaling pathway can also play a role in the development of HPE and coloboma. RA is required for proper forebrain morphogenesis and influences Shh signaling through a yet unknown mechanism. In the eye, RA is required for morphological events required for proper eye morphogenesis, and possibly for patterning eye tissues. Nodal signaling is also important, as it is required for the initiation of shh expression in the ventral forebrain. In the eye, Bmp signaling is required for eye patterning, and thus closure of the optic fissure, possibly through an antagonistic signaling 398 P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 relationship with Shh. It is thus clear that alteration of genetic pathways that subsequently influence Shh signaling, in addition to the Shh pathway itself, can result in the generation of HPE or coloboma. [26] References [27] [1] E. Matsunaga, K. Shiota, Holoprosencephaly in human embryos: epidemiologic studies of 150 cases, Teratology 16 (1977) 261–272. [2] M.M. Cohen Jr., Problems in the definition of holoprosencephaly, Am. J. Med. Genet. 103 (2001) 183–187. [3] W. DeMyer, W. Zeman, C. Palmer, The face predicts the brain: diagnostic significance of median facial anomalies for holoprosencephaly (arhinencephaly), Pediatrics 34 (1964) 256–263. [4] M. Hayhurst, S.K. McConnell, Mouse models of holoprosencephaly, Curr. Opin. Neurol. 16 (2003) 135–141. [5] M. Muenke, P.A. Beachy, Genetics of ventral forebrain development and holoprosencephaly, Curr. Opin. Genet. Dev. 10 (2000) 262–269. [6] M. Muenke, M.M.J. Cohen, Genetic approaches to understanding brain development: holoprosencephaly as a model, Men. Retard. Dev. Disabil. Res. Rev. 6 (2000) 15–21. [7] D. Wallis, M. Muenke, Mutations in holoprosencephaly, Hum. Mutat. 16 (2000) 99–108. [8] K. Shiota, S. Yamada, M. Komada, M. Ishibashi, Embryogenesis of holoprosencephaly, Am. J. Med. Genet. 143A (2007) 3079–3087. [9] C.L. Olsen, J.P. Hughes, L.G. Youngblood, M. Sharpe-Stimac, Epidemiology of holoprosencephaly and phenotypic characteristics of affected children: New York State, 1984–1989, Am. J. Med. Genet. 73 (1997) 217–226. [10] D. Higashiyama, H. Saitsu, M. Komada, T. Takigawa, M. Ishibashi, K. Shiota, Sequential developmental changes in holoprosencephalic mouse embryos exposed to ethanol during the gastrulation period, Birth Defects Res. A Clin. Mol. Teratol. 79 (2007) 513–523. [11] K.K. Sulik, D.B. Dehart, J.M. Rogers, N. Chernoff, Teratogenicity of low doses of alltrans retinoic acid in presomite mouse embryos, Teratology 51 (1995) 398–403. [12] M.M. Cohen Jr., K.K. Sulik, Perspectives on holoprosencephaly: Part II. Central nervous system, craniofacial anatomy, syndrome commentary, diagnostic approach, and experimental studies, J. Craniofac. Genet. Dev. Biol. 12 (1992) 196–244. [13] E.J. Lammer, D.T. Chen, R.M. Hoar, N.D. Agnish, P.J. Benke, J.T. Braun, C.J. Curry, P.M. Fernhoff, A.W. Grix Jr., I.T. Lott, et al., Retinoic acid embryopathy, N. Engl. J. Med. 313 (1985) 837–841. [14] S. Brown, P. Jackey, J. Lien, D. Warburton, A small deletion in the putative ‘critical region’ in chromosome 13q32 in a fetus with isolated holoprosencephaly, Am. J. Hum. Genet. 57 (1995) 606. [15] F. Gurrieri, B.J. Trask, G. van den Engh, C.M. Krauss, A. Schinzel, M.J. Pettenati, D. Schindler, J. Dietz-Band, G. Vergnaud, S.W. Scherer, et al., Physical mapping of the holoprosencephaly critical region on chromosome 7q36, Nat. Genet. 3 (1993) 247–251. [16] M. Muenke, L.J. Bone, H.F. Mitchell, I. Hart, K. Walton, K. Hall-Johnson, E.F. Ippel, J. Dietz-Band, K. Kvaloy, C.M. Fan, et al., Physical mapping of the holoprosencephaly critical region in 21q22.3, exclusion of SIM2 as a candidate gene for holoprosencephaly, and mapping of SIM2 to a region of chromosome 21 important for Down syndrome, Am. J. Hum. Genet. 57 (1995) 1074–1079. [17] J. Overhauser, H.F. Mitchell, E.H. Zackai, D.B. Tick, K. Rojas, M. Muenke, Physical mapping of the holoprosencephaly critical region in 18p11.3, Am. J. Hum. Genet. 57 (1995) 1080–1085. [18] E. Roessler, E. Belloni, K. Gaudenz, P. Jay, P. Berta, S.W. Scherer, L.C. Tsui, M. Muenke, Mutations in the human Sonic Hedgehog gene cause holoprosencephaly, Nat. Genet. 14 (1996) 357–360. [19] E. Roessler, E. Belloni, K. Gaudenz, F. Vargas, S.W. Scherer, L.C. Tsui, M. Muenke, Mutations in the C-terminal domain of Sonic Hedgehog cause holoprosencephaly, Hum. Mol. Genet. 6 (1997) 1847–1853. [20] U. Schell, J. Wienberg, A. Kohler, P. Bray-Ward, D.E. Ward, W.G. Wilson, W.P. Allen, R.R. Lebel, J.R. Sawyer, P.L. Campbell, D.J. Aughton, H.H. Punnett, E.J. Lammer, F.T. Kao, D.C. Ward, M. Muenke, Molecular characterization of breakpoints in patients with holoprosencephaly and definition of the HPE2 critical region 2p21, Hum. Mol. Genet. 5 (1996) 223–229. [21] C. Bendavid, C. Dubourg, I. Gicquel, L. Pasquier, P. Saugier-Veber, M.R. Durou, S. Jaillard, T. Frebourg, B.R. Haddad, C. Henry, S. Odent, V. David, Molecular evaluation of foetuses with holoprosencephaly shows high incidence of microdeletions in the HPE genes, Hum. Genet. 119 (2006) 1–8. [22] E. Roessler, M.V. Ouspenskaia, J.D. Karkera, J.I. Velez, A. Kantipong, F. Lacbawan, P. Bowers, J.W. Belmont, J.A. Towbin, E. Goldmuntz, B. Feldman, M. Muenke, Reduced NODAL signaling strength via mutation of several pathway members including FOXH1 is linked to human heart defects and holoprosencephaly, Am. J. Hum. Genet. 83 (2008) 18–29. [23] J.M. de la Cruz, R.N. Bamford, R.D. Burdine, E. Roessler, A.J. Barkovich, D. Donnai, A.F. Schier, M. Muenke, A loss-of-function mutation in the CFC domain of TDGF1 is associated with human forebrain defects, Hum. Genet. 110 (2002) 422–428. [24] X. Geng, C. Speirs, O. Lagutin, A. Inbal, W. Liu, L. Solnica-Krezel, Y. Jeong, D.J. Epstein, G. Oliver, Haploinsufficiency of Six3 fails to activate Sonic hedgehog expression in the ventral forebrain and causes holoprosencephaly, Dev. Cell 15 (2008) 236–247. [25] L. Nanni, J.E. Ming, M. Bocian, K. Steinhaus, D.W. Bianchi, C. Die-Smulders, A. Giannotti, K. Imaizumi, K.L. Jones, M.D. Campo, R.A. Martin, P. Meinecke, M.E. Pierpont, N.H. Robin, I.D. Young, E. Roessler, M. Muenke, The mutational [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] spectrum of the sonic hedgehog gene in holoprosencephaly: SHH mutations cause a significant proportion of autosomal dominant holoprosencephaly, Hum. Mol. Genet. 8 (1999) 2479–2488. E. Roessler, Y. Ma, M.V. Ouspenskaia, F. Lacbawan, C. Bendavid, C. Dubourg, P.A. Beachy, M. Muenke, Truncating loss-of-function mutations of DISP1 contribute to holoprosencephaly-like microform features in humans, Hum. Genet. 125 (2009) 393–400. L.G. Shaffer, A. Theisen, B.A. Bejjani, B.C. Ballif, A.S. Aylsworth, C. Lim, M. McDonald, J.W. Ellison, D. Kostiner, S. Saitta, T. Shaikh, The discovery of microdeletion syndromes in the post-genomic era: review of the methodology and characterization of a new 1q41q42 microdeletion syndrome, Genet. Med. 9 (2007) 607–616. L. Bartholin, S.E. Powers, T.A. Melhuish, S. Lasse, M. Weinstein, D. Wotton, TGIF inhibits retinoid signaling, Mol. Cell. Biol. 26 (2006) 990–1001. E. Bertolino, B. Reimund, D. Wildt-Perinic, R.G. Clerc, A novel homeobox protein which recognizes a TGT core and functionally interferes with a retinoidresponsive motif, J. Biol. Chem. 270 (1995) 31178–31188. P.A. Gongal, A.J. Waskiewicz, Zebrafish model of holoprosencephaly demonstrates a key role for TGIF in regulating retinoic acid metabolism, Hum. Mol. Genet. 17 (2008) 525–538. D. Maurus, W.A. Harris, Zic-associated holoprosencephaly: zebrafish Zic1 controls midline formation and forebrain patterning by regulating Nodal, Hedgehog, and retinoic acid signaling, Genes Dev. 23 (2009) 1461–1473. M. Brand, C.P. Heisenberg, R.M. Warga, F. Pelegri, R.O. Karlstrom, D. Beuchle, A. Picker, Y.J. Jiang, M. Furutani-Seiki, F.J. van Eeden, M. Granato, P. Haffter, M. Hammerschmidt, D.A. Kane, R.N. Kelsh, M.C. Mullins, J. Odenthal, C. NussleinVolhard, Mutations affecting development of the midline and general body shape during zebrafish embryogenesis, Development 123 (1996) 129–142. B. Feldman, M.A. Gates, E.S. Egan, S.T. Dougan, G. Rennebeck, H.I. Sirotkin, A.F. Schier, W.S. Talbot, Zebrafish organizer development and germ-layer formation require nodal-related signals, Nature 395 (1998) 181–185. K. Hatta, C.B. Kimmel, R.K. Ho, Charline Walker, The cyclops mutation blocks specification of the floor plate of the zebrafish central nervous system, Nature 350 (1991) 339–341. M. Nomura, E. Li, Smad2 role in mesoderm formation, left–right patterning and craniofacial development, Nature 393 (1998) 786–790. A.F. Schier, S.C. Neuhauss, K.A. Helde, W.S. Talbot, W. Driever, The one-eyed pinhead gene functions in mesoderm and endoderm formation in zebrafish and interacts with no tail, Development 124 (1997) 327–342. A.F. Schier, W.S. Talbot, Molecular genetics of axis formation in zebrafish, Annu. Rev. Genet. 39 (2005) 561–613. K. Grittsman, W.S. Talbot, A. Schier, Nodal signalling patterns the organizer, Development 127 (2000) 921–932. A. Mehra, J.L. Wrana, TGF-beta and the Smad signal transduction pathway, Biochem. Cell Biol. 80 (2002) 605–622. K. Gritsman, J. Zhang, S. Cheng, E. Heckscher, W.S. Talbot, A.F. Schier, The EGF-CFC protein one-eyed pinhead is essential for nodal signaling, Cell 97 (1999) 121–132. E. Reissmann, H. Jornvall, A. Blokzijl, O. Andersson, C. Chang, G. Minchiotti, M.G. Persico, C.F. Ibanez, A.H. Brivanlou, The orphan receptor ALK7 and the Activin receptor ALK4 mediate signaling by Nodal proteins during vertebrate development, Genes Dev. 15 (2001) 2010–2022. C. Yeo, M. Whitman, Nodal signals to Smads through Cripto-dependent and Cripto-independent mechanisms, Mol. Cell 7 (2001) 949–957. J. Zhang, W.S. Talbot, A.F. Schier, Positional cloning identifies zebrafish one-eyed pinhead as a permissive EGF-related ligand required during gastrulation, Cell 92 (1998) 241–251. H.M. Pogoda, L. Solnica-Krezel, W. Driever, D. Meyer, The zebrafish forkhead transcription factor FoxH1/Fast1 is a modulator of nodal signaling required for organizer formation, Curr. Biol. 10 (2000) 1041–1049. H.I. Sirotkin, M.A. Gates, P.D. Kelly, A.F. Schier, W.S. Talbot, Fast1 is required for the development of dorsal axial structures in zebrafish, Curr. Biol. 10 (2000) 1051–1054. K. Gritsman, W.S. Talbot, A.F. Schier, Nodal signaling patterns the organizer, Development 127 (2000) 921–932. F. Muller, S. Albert, P. Blader, N. Fischer, M. Hallonet, U. Strahle, Direct action of the nodal-related signal cyclops in induction of sonic hedgehog in the ventral midline of the CNS, Development 127 (2000) 3889–3897. M. Brand, C.-P. Heisenberg, R.M. Warga, F. Pelegri, R.O. Karlstrom, D. Beuchle, A. Picker, Y.-J. Jiang, M. Furutani-Seiki, F.J.M.v. Eeden, M. Granato, P. Haffter, M. Hammerschmidt, D.A. Kane, R.N. Kelsh, M.C. Mullins, J.r. Odenthal, C. NussleinVolhard, Mutations affecting development of the midline and general body shape during zebrafish embryogenesis, Development 123 (1996). A.F. Schier, S.C. Neuhauss, M. Harvey, J. Malicki, L. Solnica-Krezel, D.Y. Stainier, F. Zwartkruis, S. Abdelilah, D.L. Stemple, Z. Rangini, H. Yang, W. Driever, Mutations affecting the development of the embryonic zebrafish brain, Development 123 (1996) 165–178. G. Minchiotti, S. Parisi, G. Liguori, M. Signore, G. Lania, E.D. Adamson, C.T. Lago, M.G. Persico, Membrane-anchorage of Cripto protein by glycosylphosphatidylinositol and its distribution during early mouse development, Mech. Dev. 90 (2000) 133–142. M. Whitman, Nodal signaling in early vertebrate embryos: themes and variations, Dev. Cell 1 (2001) 605–617. P.S. Kunwar, S. Zimmerman, J.T. Bennett, Y. Chen, M. Whitman, A.F. Schier, Mixer/Bon and FoxH1/Sur have overlapping and divergent roles in Nodal signaling and mesendoderm induction, Development 130 (2003) 5589–5599. P.A. Hoodless, M. Pye, C. Chazaud, E. Labbe, L. Attisano, J. Rossant, J.L. Wrana, FoxH1 (Fast) functions to specify the anterior primitive streak in the mouse, Genes Dev. 15 (2001) 1257–1271. P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 [54] H.J. Carp, Recurrent miscarriage: genetic factors and assessment of the embryo, Isr. Med. Assoc. J. 10 (2008) 229–231. [55] P.W. Ingham, M. Placzek, Orchestrating ontogenesis: variations on a theme by sonic hedgehog, Nat. Rev. Genet. 7 (2006) 841–850. [56] S.F. Farzan, S. Singh, N.S. Schilling, D.J. Robbins, The adventures of sonic hedgehog in development and repair. III. Hedgehog processing and biological activity, Am. J. Physiol. Gastrointest. Liver Physiol. 294 (2008) G844–849. [57] Z. Chamoun, R.K. Mann, D. Nellen, D.P. von Kessler, M. Bellotto, P.A. Beachy, K. Basler, Skinny hedgehog, an acyltransferase required for palmitoylation and activity of the hedgehog signal, Science 293 (2001) 2080–2084. [58] P. Aza-Blanc, T.B. Kornberg, Ci: a complex transducer of the hedgehog signal, Trends Genet. 15 (1999) 458–462. [59] E. Roessler, K.B. El-Jaick, C. Dubourg, J.I. Velez, B.D. Solomon, D.E. Pineda-Alvarez, F. Lacbawan, N. Zhou, M. Ouspenskaia, A. Paulussen, H.J. Smeets, U. Hehr, C. Bendavid, S. Bale, S. Odent, V. David, M. Muenke, The mutational spectrum of holoprosencephaly-associated changes within the SHH gene in humans predicts loss-of-function through either key structural alterations of the ligand or its altered synthesis, Hum. Mutat. (2009). [60] X. Feng, E.G. Adiarte, S.H. Devoto, Hedgehog acts directly on the zebrafish dermomyotome to promote myogenic differentiation, Dev. Biol. 300 (2006) 736–746. [61] A. Nasevicius, S.C. Ekker, Effective targeted gene ‘knockdown’ in zebrafish, Nat. Genet. 26 (2000) 216–220. [62] H.E. Schauerte, F.J. van Eeden, C. Fricke, J. Odenthal, U. Strahle, P. Haffter, Sonic hedgehog is not required for the induction of medial floor plate cells in the zebrafish, Development 125 (1998) 2983–2993. [63] Z.M. Varga, A. Amores, K.E. Lewis, Y.L. Yan, J.H. Postlethwait, J.S. Eisen, M. Westerfield, Zebrafish smoothened functions in ventral neural tube specification and axon tract formation, Development 128 (2001) 3497–3509. [64] J.K. Chen, J. Taipale, M.K. Cooper, P.A. Beachy, Inhibition of Hedgehog signaling by direct binding of cyclopamine to Smoothened, Genes Dev. 16 (2002) 2743–2748. [65] Y. Chen, G. Struhl, Dual roles for patched in sequestering and transducing Hedgehog, Cell 87 (1996) 553–563. [66] L.V. Goodrich, D. Jung, K.M. Higgins, M.P. Scott, Overexpression of ptc1 inhibits induction of Shh target genes and prevents normal patterning in the neural tube, Dev. Biol. 211 (1999) 323–334. [67] P.W. Ingham, A.M. Taylor, Y. Nakano, Role of the Drosophila patched gene in positional signalling, Nature 353 (1991) 184–187. [68] M.J. Koudijs, M.J. den Broeder, E. Groot, F.J. van Eeden, Genetic analysis of the two zebrafish patched homologues identifies novel roles for the hedgehog signaling pathway, BMC Dev. Biol. 8 (2008) 15. [69] J.L. Mullor, I. Guerrero, A gain-of-function mutant of patched dissects different responses to the hedgehog gradient, Dev. Biol. 228 (2000) 211–224. [70] J.E. Ming, M.E. Kaupas, E. Roessler, H.G. Brunner, M. Golabi, M. Tekin, R.F. Stratton, E. Sujansky, S.J. Bale, M. Muenke, Mutations in PATCHED-1, the receptor for SONIC HEDGEHOG, are associated with holoprosencephaly, Hum. Genet. 110 (2002) 297–301. [71] F.J. van Eeden, M. Granato, U. Schach, M. Brand, M. Furutani-Seiki, P. Haffter, M. Hammerschmidt, C.P. Heisenberg, Y.J. Jiang, D.A. Kane, R.N. Kelsh, M.C. Mullins, J. Odenthal, R.M. Warga, M.L. Allende, E.S. Weinberg, C. Nusslein-Volhard, Mutations affecting somite formation and patterning in the zebrafish, Danio rerio, Development 123 (1996) 153–164. [72] R. Burke, D. Nellen, M. Bellotto, E. Hafen, K.A. Senti, B.J. Dickson, K. Basler, Dispatched, a novel sterol-sensing domain protein dedicated to the release of cholesterol-modified hedgehog from signaling cells, Cell 99 (1999) 803–815. [73] T. Caspary, M.J. Garcia-Garcia, D. Huangfu, J.T. Eggenschwiler, M.R. Wyler, A.S. Rakeman, H.L. Alcorn, K.V. Anderson, Mouse Dispatched homolog1 is required for long-range, but not juxtacrine, Hh signaling, Curr. Biol. 12 (2002) 1628–1632. [74] T. Kawakami, T. Kawcak, Y.J. Li, W. Zhang, Y. Hu, P.T. Chuang, Mouse dispatched mutants fail to distribute hedgehog proteins and are defective in hedgehog signaling, Development 129 (2002) 5753–5765. [75] Y. Ma, A. Erkner, R. Gong, S. Yao, J. Taipale, K. Basler, P.A. Beachy, Hedgehogmediated patterning of the mammalian embryo requires transporter-like function of dispatched, Cell 111 (2002) 63–75. [76] Y. Nakano, H.R. Kim, A. Kawakami, S. Roy, A.F. Schier, P.W. Ingham, Inactivation of dispatched 1 by the chameleon mutation disrupts Hedgehog signalling in the zebrafish embryo, Dev. Biol. 269 (2004) 381–392. [77] A. Inbal, S.H. Kim, J. Shin, L. Solnica-Krezel, Six3 represses nodal activity to establish early brain asymmetry in zebrafish, Neuron 55 (2007) 407–415. [78] S. Domene, E. Roessler, K.B. El-Jaick, M. Snir, J.L. Brown, J.I. Velez, S. Bale, F. Lacbawan, M. Muenke, B. Feldman, Mutations in the human SIX3 gene in holoprosencephaly are loss of function, Hum. Mol. Genet. 17 (2008) 3919–3928. [79] Y. Jeong, F.C. Leskow, K. El-Jaick, E. Roessler, M. Muenke, A. Yocum, C. Dubourg, X. Li, X. Geng, G. Oliver, D.J. Epstein, Regulation of a remote Shh forebrain enhancer by the Six3 homeoprotein, Nat. Genet. 40 (2008) 1348–1353. [80] L.L. Sandell, B.W. Sanderson, G. Moiseyev, T. Johnson, A. Mushegian, K. Young, J.P. Rey, J.X. Ma, K. Staehling-Hampton, P.A. Trainor, RDH10 is essential for synthesis of embryonic retinoic acid and is required for limb, craniofacial, and organ development, Genes Dev. 21 (2007) 1113–1124. [81] M. Maden, Retinoid signalling in the development of the central nervous system, Nat. Rev. Neurosci. 3 (2002) 843–853. [82] G. Begemann, T.F. Schilling, G.J. Rauch, R. Geisler, P.W. Ingham, The zebrafish neckless mutation reveals a requirement for raldh2 in mesodermal signals that pattern the hindbrain, Development 128 (2001) 3081–3094. [83] H. Grandel, K. Lun, G.J. Rauch, M. Rhinn, T. Piotrowski, C. Houart, P. Sordino, A.M. Kuchler, S. Schulte-Merker, R. Geisler, N. Holder, S.W. Wilson, M. Brand, Retinoic [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] [96] [97] [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [110] [111] [112] 399 acid signalling in the zebrafish embryo is necessary during pre-segmentation stages to pattern the anterior–posterior axis of the CNS and to induce a pectoral fin bud, Development 129 (2002) 2851–2865. S. Pittlik, S. Domingues, A. Meyer, G. Begemann, Expression of zebrafish aldh1a3 (raldh3) and absence of aldh1a1 in teleosts, Gene Expr. Patterns 8 (2008) 141–147. D. Chambers, L. Wilson, M. Maden, A. Lumsden, RALDH-independent generation of retinoic acid during vertebrate embryogenesis by CYP1B1, Development 134 (2007) 1369–1383. R.T. Libby, R.S. Smith, O.V. Savinova, A. Zabaleta, J.E. Martin, F.J. Gonzalez, S.W. John, Modification of ocular defects in mouse developmental glaucoma models by tyrosinase, Science 299 (2003) 1578–1581. I. Stoilov, A.N. Akarsu, I. Alozie, A. Child, M. Barsoum-Homsy, M.E. Turacli, M. Or, R.A. Lewis, N. Ozdemir, G. Brice, S.G. Aktan, L. Chevrette, M. Coca-Prados, M. Sarfarazi, Sequence analysis and homology modeling suggest that primary congenital glaucoma on 2p21 results from mutations disrupting either the hinge region or the conserved core structures of cytochrome P4501B1, Am. J. Hum. Genet. 62 (1998) 573–584. I. Stoilov, A.N. Akarsu, M. Sarfarazi, Identification of three different truncating mutations in cytochrome P4501B1 (CYP1B1) as the principal cause of primary congenital glaucoma (Buphthalmos) in families linked to the GLC3A locus on chromosome 2p21, Hum. Mol. Genet. 6 (1997) 641–647. L. Maves, C.B. Kimmel, Dynamic and sequential patterning of the zebrafish posterior hindbrain by retinoic acid, Dev. Biol. 285 (2005) 593–605. K. Niederreither, J. Vermot, N. Messaddeq, B. Schuhbaur, P. Chambon, P. Dolle, Embryonic retinoic acid synthesis is essential for heart morphogenesis in the mouse, Development 128 (2001) 1019–1031. K. Niederreither, J. Vermot, B. Schuhbaur, P. Chambon, P. Dolle, Retinoic acid synthesis and hindbrain patterning in the mouse embryo, Development 127 (2000) 75–85. R.J. White, T.F. Schilling, How degrading: Cyp26s in hindbrain development, Dev. Dyn. 237 (2008) 2775–2790. T. Kudoh, S.W. Wilson, I.B. Dawid, Distinct roles for Fgf, Wnt and retinoic acid in posteriorizing the neural ectoderm, Development 129 (2002) 4335–4346. X. Gu, F. Xu, X. Wang, X. Gao, Q. Zhao, Molecular cloning and expression of a novel CYP26 gene (cyp26d1) during zebrafish early development, Gene Expr. Patterns 5 (2005) 733–739. Q. Zhao, B. Dobbs-McAuliffe, E. Linney, Expression of cyp26b1 during zebrafish early development, Gene Expr. Patterns 5 (2005) 363–369. R.E. Hernandez, A.P. Putzke, J.P. Myers, L. Margaretha, C.B. Moens, Cyp26 enzymes generate the retinoic acid response pattern necessary for hindbrain development, Development 134 (2007) 177–187. R.J. White, Q. Nie, A.D. Lander, T.F. Schilling, Complex regulation of cyp26a1 creates a robust retinoic acid gradient in the zebrafish embryo, PLoS Biol. 5 (2007) e304. S.Q. Cohlan, Congenital anomalies in the rat produced by excessive intake of vitamin A during pregnancy, Pediatrics 13 (1954) 556–567. Y.L. Yan, T. Jowett, J.H. Postlethwait, Ectopic expression of hoxb2 after retinoic acid treatment or mRNA injection: disruption of hindbrain and craniofacial morphogenesis in zebrafish embryos, Dev. Dyn. 213 (1998) 370–385. S. Abu-Abed, P. Dolle, D. Metzger, B. Beckett, P. Chambon, M. Petkovich, The retinoic acid-metabolizing enzyme, CYP26A1, is essential for normal hindbrain patterning, vertebral identity, and development of posterior structures, Genes Dev. 15 (2001) 226–240. Y. Emoto, H. Wada, H. Okamoto, A. Kudo, Y. Imai, Retinoic acid-metabolizing enzyme Cyp26a1 is essential for determining territories of hindbrain and spinal cord in zebrafish, Dev. Biol. 278 (2005) 415–427. C. Roberts, S. Ivins, A.C. Cook, A. Baldini, P.J. Scambler, Cyp26 genes a1, b1 and c1 are down-regulated in Tbx1 null mice and inhibition of Cyp26 enzyme function produces a phenocopy of DiGeorge syndrome in the chick, Hum. Mol. Genet. 15 (2006) 3394–3410. V.S. Aggarwal, B.E. Morrow, Genetic modifiers of the physical malformations in velo-cardio-facial syndrome/DiGeorge syndrome, Dev. Disabil. Res. Rev. 14 (2008) 19–25. S.C. Butts, The facial phenotype of the velo-cardio-facial syndrome, Int. J. Pediatr. Otorhinolaryngol. 73 (2009) 343–350. F. Hale, Pigs born without eyeballs, J. Heredity 24 (1933) 105–106. A. Halilagic, M.H. Zile, M. Studer, A novel role for retinoids in patterning the avian forebrain during presomite stages, Development 130 (2003) 2039–2050. M. Maden, E. Gale, I. Kostetskii, M. Zile, Vitamin A-deficient quail embryos have half a hindbrain and other neural defects, Curr. Biol. 6 (1996) 417–426. V. Ribes, Z. Wang, P. Dolle, K. Niederreither, Retinaldehyde dehydrogenase 2 (RALDH2)-mediated retinoic acid synthesis regulates early mouse embryonic forebrain development by controlling FGF and sonic hedgehog signaling, Development 133 (2006) 351–361. D. Wotton, R.S. Lo, S. Lee, J. Massague, A Smad transcriptional corepressor, Cell 97 (1999) 29–39. K.W. Gripp, D. Wotton, M.C. Edwards, E. Roessler, L. Ades, P. Meinecke, A. Richieri-Costa, E.H. Zackai, J. Massague, M. Muenke, S.J. Elledge, Mutations in TGIF cause holoprosencephaly and link NODAL signalling to human neural axis determination, Nat. Genet. 25 (2000) 205–208. M.M. Cohen Jr., K. Shiota, Teratogenesis of holoprosencephaly, Am. J. Med. Genet. 109 (2002) 1–15. C. Dubourg, L. Lazaro, L. Pasquier, C. Bendavid, M. Blayau, F. Le Duff, M.R. Durou, S. Odent, V. David, Molecular screening of SHH, ZIC2, SIX3, and TGIF genes in patients with features of holoprosencephaly spectrum: mutation review and genotype–phenotype correlations, Hum. Mutat. 24 (2004) 43–51. 400 P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 [113] E. Roessler, M. Muenke, Holoprosencephaly: a paradigm for the complex genetics of brain development, J. Inherit. Metab. Dis. 21 (1998) 481–497. [114] J.Z. Jin, S. Gu, P. McKinney, J. Ding, Expression and functional analysis of Tgif during mouse midline development, Dev. Dyn. 235 (2006) 547–553. [115] L. Mar, P.A. Hoodless, Embryonic fibroblasts from mice lacking Tgif were defective in cell cycling, Mol. Cell. Biol. 26 (2006) 4302–4310. [116] J. Shen, C.A. Walsh, Targeted disruption of Tgif, the mouse ortholog of a human holoprosencephaly gene, does not result in holoprosencephaly in mice, Mol. Cell. Biol. 25 (2005) 3639–3647. [117] C. Kuang, Y. Xiao, L. Yang, Q. Chen, Z. Wang, S.J. Conway, Y. Chen, Intragenic deletion of Tgif causes defects in brain development, Hum. Mol. Genet. 15 (2006) 3508–3519. [118] J.J. Sanz-Ezquerro, C. Tickle, Autoregulation of Shh expression and Shh induction of cell death suggest a mechanism for modulating polarising activity during chick limb development, Development 127 (2000) 4811–4823. [119] R.A. Schneider, D. Hu, J.L. Rubenstein, M. Maden, J.A. Helms, Local retinoid signaling coordinates forebrain and facial morphogenesis by maintaining FGF8 and SHH, Development 128 (2001) 2755–2767. [120] V. Ribes, I. Le Roux, M. Rhinn, B. Schuhbaur, P. Dolle, Early mouse caudal development relies on crosstalk between retinoic acid, Shh and Fgf signalling pathways, Development 136 (2009) 665–676. [121] M.H. Chen, Y.J. Li, T. Kawakami, S.M. Xu, P.T. Chuang, Palmitoylation is required for the production of a soluble multimeric Hedgehog protein complex and longrange signaling in vertebrates, Genes Dev. 18 (2004) 641–659. [122] J. Feng, B. White, O.V. Tyurina, B. Guner, T. Larson, H.Y. Lee, R.O. Karlstrom, J.D. Kohtz, Synergistic and antagonistic roles of the Sonic hedgehog N- and Cterminal lipids, Development 131 (2004) 4357–4370. [123] N.A. Sanek, Y. Grinblat, A novel role for zebrafish zic2a during forebrain development, Dev. Biol. 317 (2008) 325–335. [124] N.A. Sanek, A.A. Taylor, M.K. Nyholm, Y. Grinblat, Zebrafish zic2a patterns the forebrain through modulation of Hedgehog-activated gene expression, Development 136 (2009) 3791–3800. [125] Y.R. Barishak, Embryology of the eye and its adnexae, Dev. Ophthalmol. 24 (1992) 1–142. [126] D. Taylor, Developmental abnormalities of the optic nerve and chiasm, Eye 21 (2007) 1271–1284. [127] M. Asai-Coakwell, C.R. French, K.M. Berry, M. Ye, R. Koss, M. Somerville, R. Mueller, V. van Heyningen, A.J. Waskiewicz, O.J. Lehmann, GDF6, a novel locus for a spectrum of ocular developmental anomalies, Am. J. Hum. Genet. 80 (2007) 306–315. [128] C.Y. Gregory-Evans, M.J. Williams, S. Halford, K. Gregory-Evans, Ocular coloboma: a reassessment in the age of molecular neuroscience, J. Med. Genet. 41 (2004) 881–891. [129] E. Ferda Percin, L.A. Ploder, J.J. Yu, K. Arici, D.J. Horsford, A. Rutherford, B. Bapat, D.W. Cox, A.M. Duncan, V.I. Kalnins, A. Kocak-Altintas, J.C. Sowden, E. Traboulsi, M. Sarfarazi, R.R. McInnes, Human microphthalmia associated with mutations in the retinal homeobox gene CHX10, Nat. Genet. 25 (2000) 397–401. [130] D. Morrison, D. FitzPatrick, I. Hanson, K. Williamson, V. van Heyningen, B. Fleck, I. Jones, J. Chalmers, H. Campbell, National study of microphthalmia, anophthalmia, and coloboma (MAC) in Scotland: investigation of genetic aetiology, J. Med. Genet. 39 (2002) 16–22. [131] R.V. Jamieson, R. Perveen, B. Kerr, M. Carette, J. Yardley, E. Heon, M.G. Wirth, V. van Heyningen, D. Donnai, F. Munier, G.C. Black, Domain disruption and mutation of the bZIP transcription factor, MAF, associated with cataract, ocular anterior segment dysgenesis and coloboma, Hum. Mol. Genet. 11 (2002) 33–42. [132] L.E. Vissers, C.M. van Ravenswaaij, R. Admiraal, J.A. Hurst, B.B. de Vries, I.M. Janssen, W.A. van der Vliet, E.H. Huys, P.J. de Jong, B.C. Hamel, E.F. Schoenmakers, H.G. Brunner, J.A. Veltman, A.G. van Kessel, Mutations in a new member of the chromodomain gene family cause CHARGE syndrome, Nat. Genet. 36 (2004) 955–957. [133] M. Torres, E. Gomez-Pardo, P. Gruss, Pax2 contributes to inner ear patterning and optic nerve trajectory, Development 122 (1996) 3381–3391. [134] M.R. Eccles, L.A. Schimmenti, Renal-coloboma syndrome: a multi-system developmental disorder caused by PAX2 mutations, Clin. Genet. 56 (1999) 1–9. [135] L.A. Schimmenti, J. de la Cruz, R.A. Lewis, J.D. Karkera, G.S. Manligas, E. Roessler, M. Muenke, Novel mutation in sonic hedgehog in non-syndromic colobomatous microphthalmia, Am. J. Med. Genet. 116A (2003) 215–221. [136] S.J. Hornby, S.J. Ward, C.E. Gilbert, L. Dandona, A. Foster, R.B. Jones, Environmental risk factors in congenital malformations of the eye, Ann. Trop. Paediatr. 22 (2002) 67–77. [137] J.G. Wilson, C.B. Roth, J. Warkany, An analysis of the syndrome of malformations induced by maternal vitamin A deficiency. Effects of restoration of vitamin A at various times during gestation, Am. J. Anat. 92 (1953) 189–217. [138] R. Macdonald, K.A. Barth, Q. Xu, N. Holder, I. Mikkola, S.W. Wilson, Midline signalling is required for Pax gene regulation and patterning of the eyes, Development 121 (1995) 3267–3278. [139] C.J. Neumann, C. Nuesslein-Volhard, Patterning of the zebrafish retina by a wave of sonic hedgehog activity, Science 289 (2000) 2137–2139. [140] M. Take-uchi, J.D. Clarke, S.W. Wilson, Hedgehog signalling maintains the optic stalk-retinal interface through the regulation of Vax gene activity, Development 130 (2003) 955–968. [141] S.C. Ekker, A.R. Ungar, P. Greenstein, D.P. von Kessler, J.A. Porter, R.T. Moon, P.A. Beachy, Patterning activities of vertebrate hedgehog proteins in the developing eye and brain, Curr. Biol. 5 (1995) 944–955. [142] L.A. Schimmenti, G.S. Manligas, P.A. Sieving, Optic nerve dysplasia and renal insufficiency in a family with a novel PAX2 mutation, Arg115X: further [143] [144] [145] [146] [147] [148] [149] [150] [151] [152] [153] [154] [155] [156] [157] [158] [159] [160] [161] [162] [163] [164] [165] [166] [167] [168] ophthalmologic delineation of the renal-coloboma syndrome, Ophthalmic Genet. 24 (2003) 191–202. A. Taranta, A. Palma, V. De Luca, A. Romanzo, L. Massella, F. Emma, L. Dello Strologo, Renal-coloboma syndrome: a single nucleotide deletion in the PAX2 gene at Exon 8 is associated with a highly variable phenotype, Clin. Nephrol. 67 (2007) 1–4. P. Elms, P. Siggers, D. Napper, A. Greenfield, R. Arkell, Zic2 is required for neural crest formation and hindbrain patterning during mouse development, Dev. Biol. 264 (2003) 391–406. E. Herrera, L. Brown, J. Aruga, R.A. Rachel, G. Dolen, K. Mikoshiba, S. Brown, C.A. Mason, Zic2 patterns binocular vision by specifying the uncrossed retinal projection, Cell 114 (2003) 545–557. J. Lee, J.R. Willer, G.B. Willer, K. Smith, R.G. Gregg, J.M. Gross, Zebrafish blowout provides genetic evidence for Patched1-mediated negative regulation of Hedgehog signaling within the proximal optic vesicle of the vertebrate eye, Dev. Biol. 319 (2008) 10–22. T.H. Kim, J. Goodman, K.V. Anderson, L. Niswander, Phactr4 regulates neural tube and optic fissure closure by controlling PP1-, Rb-, and E2F1-regulated cell-cycle progression, Dev. Cell 13 (2007) 87–102. V.A. Ruzhynsky, M. Furimsky, D.S. Park, V.A. Wallace, R.S. Slack, E2F4 is required for early eye patterning, Dev. Neurosci. 31 (2009) 238–246. M.W. Seeliger, H.K. Biesalski, B. Wissinger, H. Gollnick, S. Gielen, J. Frank, S. Beck, E. Zrenner, Phenotype in retinol deficiency due to a hereditary defect in retinol binding protein synthesis, Invest. Ophthalmol. Vis. Sci. 40 (1999) 3–11. H. Ozeki, S. Shirai, Developmental eye abnormalities in mouse fetuses induced by retinoic acid, Jpn. J. Ophthalmol. 42 (1998) 162–167. D.L. Stull, K.C. Wikler, Retinoid-dependent gene expression regulates early morphological events in the development of the murine retina, J. Comp. Neurol. 417 (2000) 289–298. G.A. Hyatt, E.A. Schmitt, N. Marsh-Armstrong, P. McCaffery, U.C. Drager, J.E. Dowling, Retinoic acid establishes ventral retinal characteristics, Development 122 (1996) 195–204. J.A. Helms, C.H. Kim, D. Hu, R. Minkoff, C. Thaller, G. Eichele, Sonic hedgehog participates in craniofacial morphogenesis and is down-regulated by teratogenic doses of retinoic acid, Dev. Biol. 187 (1997) 25–35. N. Marsh-Armstrong, P. McCaffery, W. Gilbert, J.E. Dowling, U.C. Drager, Retinoic acid is necessary for development of the ventral retina in zebrafish, Proc. Natl. Acad. Sci. U. S. A. 91 (1994) 7286–7290. A. Perz-Edwards, N.L. Hardison, E. Linney, Retinoic acid-mediated gene expression in transgenic reporter zebrafish, Dev. Biol. 229 (2001) 89–101. M.J. Connor, Oxidation of retinol to retinoic acid as a requirement for biological activity in mouse epidermis, Cancer Res. 48 (1988) 7038–7040. M.J. Connor, M.H. Smit, Terminal-group oxidation of retinol by mouse epidermis. Inhibition in vitro and in vivo, Biochem. J. 244 (1987) 489–492. T.J. Schuh, B.L. Hall, J.C. Kraft, M.L. Privalsky, D. Kimelman, v-erbA and citral reduce the teratogenic effects of all-trans retinoic acid and retinol, respectively, in Xenopus embryogenesis, Development 119 (1993) 785–798. A.W. See, M. Clagett-Dame, The temporal requirement for vitamin A in the developing eye: mechanism of action in optic fissure closure and new roles for the vitamin in regulating cell proliferation and adhesion in the embryonic retina, Dev. Biol. 325 (2009) 94–105. A. Molotkov, N. Molotkova, G. Duester, Retinoic acid guides eye morphogenetic movements via paracrine signaling but is unnecessary for retinal dorsoventral patterning, Development 133 (2006) 1901–1910. M. Asai-Coakwell, C.R. French, M. Ye, K. Garcha, K. Bigot, A.G. Perera, K. Staehling-Hampton, S.C. Mema, B. Chanda, A. Mushegian, S. Bamforth, M.R. Doschak, G. Li, M.B. Dobbs, P.F. Giampietro, B.P. Brooks, P. Vijayalakshmi, Y. Sauve, M. Abitbol, P. Sundaresan, V. van Heyningen, O. Pourquie, T.M. Underhill, A.J. Waskiewicz, O.J. Lehmann, Incomplete penetrance and phenotypic variability characterize Gdf6-attributable oculo-skeletal phenotypes, Hum. Mol. Genet. 18 (2009) 1110–1121. C.R. French, T. Erickson, D.V. French, D.B. Pilgrim, A.J. Waskiewicz, Gdf6a is required for the initiation of dorsal–ventral retinal patterning and lens development, Dev. Biol. 333 (2009) 37–47. N.J. Gosse, H. Baier, An essential role for Radar (Gdf6a) in inducing dorsal fate in the zebrafish retina, Proc. Natl. Acad. Sci. U. S. A. 106 (2009) 2236–2241. P. Bakrania, M. Efthymiou, J.C. Klein, A. Salt, D.J. Bunyan, A. Wyatt, C.P. Ponting, A. Martin, S. Williams, V. Lindley, J. Gilmore, M. Restori, A.G. Robson, M.M. Neveu, G.E. Holder, J.R. Collin, D.O. Robinson, P. Farndon, H. Johansen-Berg, D. Gerrelli, N.K. Ragge, Mutations in BMP4 cause eye, brain, and digit developmental anomalies: overlap between the BMP4 and hedgehog signaling pathways, Am. J. Hum. Genet. 82 (2008) 304–319. L. Zhao, H. Saitsu, X. Sun, K. Shiota, M. Ishibashi, Sonic hedgehog is involved in formation of the ventral optic cup by limiting Bmp4 expression to the dorsal domain, Mech. Dev. (2009). Y. Ohkubo, C. Chiang, J.L. Rubenstein, Coordinate regulation and synergistic actions of BMP4, SHH and FGF8 in the rostral prosencephalon regulate morphogenesis of the telencephalic and optic vesicles, Neuroscience 111 (2002) 1–17. A.M. Hever, K.A. Williamson, V. van Heyningen, Developmental malformations of the eye: the role of PAX6, SOX2 and OTX2, Clin. Genet. 69 (2006) 459–470. N.K. Ragge, A.G. Brown, C.M. Poloschek, B. Lorenz, R.A. Henderson, M.P. Clarke, I. Russell-Eggitt, A. Fielder, D. Gerrelli, J.P. Martinez-Barbera, P. Ruddle, J. Hurst, J.R. Collin, A. Salt, S.T. Cooper, P.J. Thompson, S.M. Sisodiya, K.A. Williamson, D.R. Fitzpatrick, V. van Heyningen, I.M. Hanson, Heterozygous mutations of OTX2 cause severe ocular malformations, Am. J. Hum. Genet. 76 (2005) 1008–1022. P.A. Gongal et al. / Biochimica et Biophysica Acta 1812 (2011) 390–401 [169] A. Wyatt, P. Bakrania, D.J. Bunyan, R.J. Osborne, J.A. Crolla, A. Salt, C. Ayuso, R. Newbury-Ecob, Y. Abou-Rayyah, J.R. Collin, D. Robinson, N. Ragge, Novel heterozygous OTX2 mutations and whole gene deletions in anophthalmia, microphthalmia and coloboma, Hum. Mutat. 29 (2008) E278–283. [170] N. Fujimura, M.M. Taketo, M. Mori, V. Korinek, Z. Kozmik, Spatial and temporal regulation of Wnt/beta-catenin signaling is essential for development of the retinal pigment epithelium, Dev. Biol. 334 (2009) 31–45. [171] J.R. Martinez-Morales, V. Dolez, I. Rodrigo, R. Zaccarini, L. Leconte, P. Bovolenta, S. Saule, OTX2 activates the molecular network underlying retina pigment epithelium differentiation, J. Biol. Chem. 278 (2003) 21721–21731. 401 [172] C.R. French, T. Erickson, D. Callander, K.M. Berry, R. Koss, D.W. Hagey, J. Stout, K. Wuennenberg-Stapleton, J. Ngai, C.B. Moens, A.J. Waskiewicz, Pbx homeodomain proteins pattern both the zebrafish retina and tectum, BMC Dev. Biol. 7 (2007) 85. [173] P. Wang, X. Liang, J. Yi, Q. Zhang, Novel SOX2 mutation associated with ocular coloboma in a Chinese family, Arch. Ophthalmol. 126 (2008) 709–713. [174] O.V. Taranova, S.T. Magness, B.M. Fagan, Y. Wu, N. Surzenko, S.R. Hutton, L.H. Pevny, SOX2 is a dose-dependent regulator of retinal neural progenitor competence, Genes Dev. 20 (2006) 1187–1202. Biochimica et Biophysica Acta 1812 (2011) 402–414 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Fish models in prion biology: Underwater issues☆ Edward Málaga-Trillo a,⁎,1, Evgenia Salta b,1, Antonio Figueras c, Cynthia Panagiotidis d, Theodoros Sklaviadis b,⁎ a Department of Biology, University of Konstanz, 78457 Konstanz, Germany Department of Pharmacology, Aristotle University of Thessaloniki, 54124, Thessaloniki, Greece c Instituto de Investigaciones Marinas, CSIC, 36208, Vigo, Spain d Centre for Research and Technology-Hellas, Institute of Agrobiotechnology, 57001, Thessaloniki, Greece b a r t i c l e i n f o Article history: Received 8 December 2009 Received in revised form 11 September 2010 Accepted 21 September 2010 Available online 7 October 2010 Keywords: Prion protein Transmissible spongiform encephalopathy Neurodegeneration Fish Zebrafish a b s t r a c t Transmissible spongiform encephalopathies (TSEs), otherwise known as prion disorders, are fatal diseases causing neurodegeneration in a wide range of mammalian hosts, including humans. The causative agents – prions – are thought to be composed of a rogue isoform of the endogenous prion protein (PrP). Beyond these and other basic concepts, fundamental questions in prion biology remain unanswered, such as the physiological function of PrP, the molecular mechanisms underlying prion pathogenesis, and the origin of prions. To date, the occurrence of TSEs in lower vertebrates like fish and birds has received only limited attention, despite the fact that these animals possess bona fide PrPs. Recent findings, however, have brought fish before the footlights of prion research. Fish models are beginning to provide useful insights into the roles of PrP in health and disease, as well as the potential risk of prion transmission between fish and mammals. Although still in its infancy, the use of fish models in TSE research could significantly improve our basic understanding of prion diseases, and also help anticipate risks to public health. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Introduction 1.1. Prion disorders Prion diseases or TSEs are a group of rare but fatal neurological disorders that affect humans and animals. Whether sporadic, inherited or acquired, these illnesses generally correlate with the accumulation of misfolded PrP in the brain and the appearance of widespread neurodegeneration after long incubation times [1]. The classical histopathological landmarks of TSEs are spongiform vacuolation, neuronal loss and astrocytic gliosis, whereas the main clinical manifestations in humans include progressive dementia, cerebellar ataxia and myoclonus [2]. Interestingly, although such symptoms are also observed in more common neurodegenerative disorders like Abbreviations: TSEs, transmissible spongiform encephalopathies; PrP, prion protein; BSE, bovine spongiform encephalopathy; CJD, Creutzfeldt–Jakob disease; GPI, glycosylphospatidylinositol; MBM, meat and bone meal; p.i., post inoculation; i.c., intracerebrally; PK, proteinase K; GI, gastrointestinal ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Corresponding authors. E. Málaga-Trillo is to be contacted at Department of Biology, University of Konstanz, Universitätstrasse 10, 78457, Konstanz, Germany. Tel.: +49 7531 884581; fax: +49 7531 884863. T. Sklaviadis, Department of Pharmacology, School of Pharmacy, Aristotle University of Thessaloniki (AUTH), AUTH Campus, 54214 Thessaloniki, Greece. Tel.: +30 2310 997615; fax: +30 2310 997720. E-mail addresses: [email protected] (E. Málaga-Trillo), [email protected] (T. Sklaviadis). 1 Equally contributing authors. 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.09.013 Alzheimer's disease (AD) and Parkinson's disease (PD), prion diseases have received special attention because of their infectious nature and the associated risk of epidemics. In fact, prion diseases gained considerable notoriety in the 1990s because of the massive outbreak of mad cow disease (bovine spongiform encephalopathy, BSE) in Europe and the negative impact it had on public health and the food industry. TSEs are, however, not a recent phenomenon. Their history dates back to the 1700s when scrapie, a related neurological condition among sheep and goats, was first described in England [3]. Other TSEs in animals include chronic wasting disease (CWD) in deer and elk, transmissible mink encephalopathy and feline spongiform encephalopathy (FSE) [4]. The most common TSEs in humans, Creutzfeldt–Jacob disease (CJD) and Gerstmann Sträussler Scheinker syndrome (GSS), were discovered much later (1920s and 1930s, respectively [5]), and it was not until the 1950s that the study of kuru, another human prion disease, set off a chain of revolutionary discoveries that would redefine the meaning of “pathogen” and push a biological dogma against the ropes. 1.2. The nature of the infectious TSE agent Kuru was originally described as a neurological epidemic of the Fore linguistic group in Papua New Guinea [2]. For some time, the disease was believed to be infectious, and possibly transmitted during cannibalistic rituals. Nevertheless, failure to transmit kuru to laboratory animals led to the erroneous assumption that it had a genetic basis. Towards 1959, remarkable similarities between kuru, CJD and scrapie E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 were recognized at the neuropathological, clinical and epidemiological levels [6,7]. Since scrapie had been already shown to be transmissible [8], it became apparent that kuru and CJD were also of infectious etiology. Indeed, by 1968, both human diseases had been successfully transmitted to chimpanzees [9,10]. Despite these advances, the nature of the scrapie agent continued to be the subject of intense debate for many years. Because TSEs exhibit strain variation and long incubation periods, it was argued that the agent could be a slow-virus [11]. However, direct proof for this “viral hypothesis” has been difficult to obtain [12]. Instead, it has been shown that the transmissible agent is resistant to formalin (a chemical that inactivates viruses) and to various other treatments that destroy nucleic acids, but not to proteolytic digestion [13–15]. These and other developments finally led to the enunciation of the “protein-only hypothesis” by Stanley Prusiner (1982), according to which TSEs are caused by prions, small proteinaceous particles capable of replicating in the absence of nucleic acids [14,16]. The protein-only hypothesis has not remained uncontested. While some continue to advocate the existence of an unidentified scrapie agent [17–19], others see the notion of an infectious, self-replicating protein as a clear violation of Crick's central dogma of molecular biology [20], which excludes the flow of biological information from protein to protein. Notwithstanding the controversy, Prusiner's theory reasonably accounts for the complex etiology of TSEs, the physicochemical properties of the agent, and the existence of scrapie strains [3]. Today, it is widely accepted that prions are largely – if not entirely – composed of PrPSc, an abnormally folded isoform of the cellular prion protein (PrPC) [21]. The molecular mechanism by which normal PrPC “converts” into infectious PrPSc is, however, poorly understood. It has been proposed that PrPSc can associate with PrPC and induce its conversion into more PrPSc, triggering a chain reaction [22]. Alternatively, PrPSc and PrPC may coexist in equilibrium until mutation or sporadic events favor the formation of oligomeric PrPSc seeds, which undergo repeated aggregation and fragmentation cycles [23]. In either case, the concept of PrP conversion is central to the protein-only hypothesis because it explains the ability of prions to replicate and propagate disease. 1.3. The prion protein In 1982, PrPSc was identified as the major constituent of infective fractions purified from hamster brain homogenates [24]. Subsequent characterization revealed that the pathogenic protein was host-encoded and not the product of a viral gene, as it had been assumed previously [25,26]. It was also established that PrPSc is posttranslationally derived from PrPC [27]. Thus, although the two isoforms differ greatly in their spatial conformation [28], they have the same amino acid sequence and are encoded by the same single-copy gene, Prnp [29]. Under normal conditions, PrPC is a glycoprotein tethered to the outer plasma membrane by a glycosylphosphatidylinositol (GPI) anchor. Its unique molecular structure, studied by nuclear magnetic resonance (NMR) and crystallographic techniques, can be roughly divided into two halves: a flexible N-terminal domain rich in repetitive motifs, and a C-terminal globular domain containing a characteristic array of three α-helices and two β-sheets. At the center of the polypeptide chain, a short hydrophobic stretch connects the two major domains [30,31]. In contrast, attempts to resolve the molecular structure of PrPSc have not been successful. Almost 30 years after the discovery of PrP, there is still no consensus about what its physiological role might be. Since PrP has no obvious similarities to known protein families, functional affinities cannot be inferred from amino acid sequence comparisons. Neither can organ- or tissue-specific functions be hypothesized because PrP is widely expressed in most adult and embryonic cell types [32–35]. Astonishingly, the generation and analysis of PrP knockout mice failed to provide conclusive information about the physiological role of PrP, as these animals develop and behave rather normally. Aside from a few subtle 403 abnormalities, their only clear “phenotype” is their resistance to prion infection [36]. It is probably not exaggerated to say that as many PrP functions have been proposed as methods have been employed to address this question. Thus, the list of putative PrP roles in vitro and in vivo includes activities as diverse as cytoprotection from apoptosis [37] and oxidative stress [38], olfactory behavior [39], copper metabolism [40], neurogenesis [41], neurite outgrowth and neuronal survival [42], lymphocyte activation [43], hematopoietic stem cell self-renewal [44], synaptic function [45], activation of signal transduction [46], and binding to cell adhesion molecules [47]. Unfortunately, it is not clear how many of these findings may be of actual physiological relevance. In addition, it seems unlikely that a single molecular function of PrP could account for such diverse biological roles. The clear correlation between PrP misfolding and neurodegeneration may suggest that PrPSc is the direct cause of prion disease. Nevertheless, solid experimental evidence indicates that prions cannot induce neuronal damage in the absence of PrPC. For instance, PrP knockout mice are resistant to scrapie inoculation [48], even if they carry PrPC-expressing neurografts [49], or if PrPC depletion is induced postnatally [50]. Notably, transgenic mice expressing only a GPI-anchorless version of PrP (thus unable to attach to the plasma membrane) can efficiently replicate prions without developing prion disease [51]. Altogether, these studies demonstrate that prion neurotoxicity is not triggered by PrPSc itself, but by an activity of PrPC at the cell surface. Accordingly, it has been proposed that prion replication might cause PrPC to either lose a neuroprotective role, gain a neurotoxic one, or subvert a neuroprotective role into a neurotoxic one [52]. Therefore, a thorough characterization of PrPC function will likely be instrumental in ascertaining the molecular basis of prioninduced neurodegeneration. 1.4. Open challenges in prion biology and pathogenesis The past five decades have seen remarkable achievements that shaped our current understanding of TSEs. These include the transmissibility of kuru and CJD to laboratory animals, the isolation of the scrapie agent, the development of the protein-only hypothesis, the identification of PrP and its isoforms, and the study of prion propagation in genetically modified mice. Yet, very little has been learned about basic aspects of prion pathogenesis, such as the mechanisms of PrP misfolding and prion replication, the physiological function of PrP, or the cellular pathways through which prions induce neurodegeneration. To this day, prion disorders remain incurable, and neither early diagnostic methods nor anti-prion drugs have successfully been developed. Of special concern is the possibility that animal prion diseases may be transmitted to humans via the food chain. In fact, experimental evidence indicates that transmission of BSE to humans was the likeliest cause for the appearance of a new variant form of CJD (vCJD) [53,54]. These findings illustrate the need to ensure food safety, not only through efficient quality control but also by experimentally assessing the risk of prion transmission across a wide range of animal hosts. So far, only beef and lamb have been identified as potential sources of dietary prion infections. Other major animal food sources like fish and poultry appear to have remained free of prion contamination. However, it is not clear whether these organisms are susceptible to prion pathogenesis because TSE research has been carried out predominantly in mice and other mammals. Therefore, the study of prion biology in non-mammals, particularly in fish, is of great relevance to food safety. Fish constitute a vital food resource for humans and animals worldwide (Table 1), with direct impact on important economic activities like fishing and aquaculture. In addition, several fish species have become valuable tools in biomedical research. One of them, the zebrafish (Danio rerio), is an excellent model to study vertebrate development and the biology of human disease. It has many advantages as a genetic system, such as its 404 E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 Table 1 Commercially important fish species worldwide. Source: Food and Agriculture Organization of the United Nations, Fisheries and Aquaculture Department (http://www.fao.org/fishery/en). Common name Scientific name Main producers Salmon Salmo salar Japan, Europe, North America Austria, Germany, Hungary, Poland Africa, China, Papua New Guinea, USA, China Philippines, Thailand, Indonesia Carp Tilapia Main consumers Chile, Norway, Scotland, USA, Canada Cyprinus carpio Central Europe, China Oreochromis niloticus niloticus Tuna Thunnus albacores Catfish Ictalutus punctatus Trout Oncorhynchus mykiss Seabream Sparus aurata Sea bass Dicentrarchus labrax Japan, Australia Japan, Australia USA, China USA, China Europe, USA, Chile Europe, USA, Japan Europe Europe Mediterranean countries Mediterranean countries divergent regions of the protein, respectively. Remarkably, and despite having undergone substantial sequence variation, the globular domain retained its structural fold throughout evolution [67,69]. These observations strongly suggest that the globular domain carries out an important function, which was already present in the last common ancestor of all vertebrates. Whether prion behavior was already an intrinsic attribute of such primitive PrPs, or whether it appeared late in evolution remains to be clarified. So far, TSEs have only been reported in mammals. This is not surprising, since prion biology is a field primarily focused on the study of human diseases. Nevertheless, if PrP is intrinsically able to form infectious prions, then its presence in non-mammals implies that these animals would – at least in theory – be susceptible to some form of prion disease. Evaluating this scenario is particularly difficult without basic knowledge about PrP conversion and prion replication in nonmammalian species. Moreover, it is not even clear what the eventual manifestations of TSEs in fish or birds might be. Clearly, extensive experimentation will be required to establish whether these organisms can acquire and transmit prion disease. 2.2. Prion transmission and the species barrier large offspring size and short generation time [55]. Furthermore, zebrafish embryos are transparent and develop externally, making them amenable to detailed cellular analyses and genetic manipulations. Moreover, the zebrafish is rapidly emerging as a valuable tool in clinical and pharmaceutical research aimed at drug discovery and validation [56,57]. In this article, we review current knowledge about the occurrence of PrPs, prions, and TSEs among vertebrates, focusing on recent advances using fish models to characterize the physiological function of PrP and the possible risk of prion transmission between fish and mammals. 2. Prion disease: a mammalian attribute? 2.1. PrP and prion-like “behavior” Besides PrP, a handful of mammalian proteins appear to be able to replicate in a prion-like manner. Among them are the molecular culprits of known neurodegenerative disorders, such as amyloid-β (Alzheimer's disease), α-synuclein (Parkinson's disease), tau (Tauopathies) and amyloid A (AA amyloidosis) [58]. Such observations are undoubtedly intriguing, yet it still needs to be established whether the rogue versions of these proteins are indeed infectious and, like bona fide prions, capable of spreading disease between individuals or even across species. In fungi, at least six different proteins exhibit prion-like behavior, as judged by their ability to propagate phenotypic traits between cells without the need for nucleic acids [59]. However, these proteins have neither structural nor functional similarities to PrP, or to one another. Thus, while the study of fungal prions has provided valuable information about the molecular basis of prion conversion, it is not clear to what extent this knowledge may apply to mammalian prion disease [60]. In addition to mammals, PrP homologues have also been identified in birds [61], reptiles [62], amphibians [63] and fish [64], but not in invertebrates. Comparison of amino acid sequences indicates that PrPs have evolved differently in each vertebrate class. For instance, although mammalian and avian sequences are strongly conserved, with similarity scores of ~90% in each group, conservation between the two groups is rather low (~30% similarity) [65,66]. Likewise, PrP sequences among bony fish are relatively well conserved (50–60% similarity), but exhibit only 20–25% similarity to mammalian sequences [67]. PrPs also exhibit considerable variation in size, ranging from ~260 amino acids in mammals and birds to ~600 amino acids in fish. Despite these large differences in length and sequence, all vertebrate PrPs share the same protein domains (Fig. 1) [67,68]. The degree of conservation varies significantly along the polypeptide, with the short hydrophobic stretch and the repetitive domain being the most conserved and the most Prion transmission across different species is limited by the so-called species barrier. Following primary inoculation between two species, only some animals develop disease, with long incubation periods that may reach the normal life span of the species. This phenomenon is clearly illustrated by the resistance to TSE observed in rabbits, even after intracranial inoculation with TSE agents of sheep, mouse and human origin [70]. It has been suggested that the unique structural characteristics of rabbit PrP make it less prone to pathogenic conversion [70,71]. Typically, species barriers become attenuated after serial passaging of a prion strain, resulting in shortened incubation times and the consequent adaptation of the host to the specific strain [72]. At the molecular level, the efficiency of interspecies prion transmission appears to depend on the degree of sequence and structural identity between the PrPs of the two species involved [73]. For instance, wild type mice do not become infected with hamster prions, whereas transgenic mice overexpressing hamster PrP do [74]. However, BSE can effectively be transmitted to a wide range of species and maintain its pathogenic characteristics, even when passaged through an intermediate species carrying a different PrP coding gene [75]. It has been proposed that every species harbors a number of possible PrPSc conformational variants, with a thermodynamically “preferred” conformation (strain) being the decisive factor for interspecies transmission. According to this “conformational selection model” [76], a significant overlap between the preferred PrPSc configurations of two species is likely to result in a relatively easy transmission of prions between them [77]. On the contrary, two species not sharing common PrPSc structures would have to overcome a considerable barrier to transmission, and pathogenesis would not ensue without modification of the strain type. Therefore, evaluation of the prion transmission from one species to another should also be based on the structural identity between the host and donor PrP molecules and not exclusively on their amino acid sequence [78]. The route of infection is another parameter affecting the magnitude of species barriers. Most inoculation studies rely on the use of intracranial injection because “natural” per os infection is usually less effective [79,80]. Notably, even after homologous oral transmission of BSE to cattle, animals may remain presymptomatic for 28 months, despite the evident PrPSc deposition in their brains [81]. Although the relative importance of the aforementioned factors has been extensively studied, the outcome of an intra-species prion transmission cannot be predicted in advance. It is noteworthy, that two different TSE strains that can effectively infect one species may present completely modified transmissibility potentials when transmitted to a second species. Thus, it has been proposed that a more suitable term to describe these E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 405 Fig. 1. Fish and mammalian prion proteins. Upper panel: Fish possess two prion proteins, PrP-1 and PrP-2. Despite differences in sequence and length, fish and mammalian PrPs share the same protein domains (drawn at scale). Lower panel: Recent studies indicate that a common involvement of vertebrate PrPs in cell–cell communication. In cultured cells, PrPs elicit cell–cell contact formation. In the zebrafish embryo, PrP-1 and PrP-2 are deployed during gastrulation and neural development, respectively. L = leader peptide, R = repetitive domains, H = hydrophobic stretch, G = globular domain; orange triangles indicate the ends of the mature polypeptides. PrP localization at cell contacts is depicted in red. phenomena would be “transmission barriers” instead of “species barriers” [76]. Recent data have led to the reconsideration of many parameters that were thought to be indicative of a species barrier, such as the lack of clinical symptoms. Interestingly, cases of subclinical infection in mammals displaying neuropathological and biochemical abnormalities, often including PrPSc accumulation, have been reported [82]. Thus, prion-inoculated animals harboring high infectivity levels, and even abnormal PrP aggregates in their brains, may under certain circumstances remain asymptomatic carriers throughout their lifetime. Notably, subclinical disease has been reported even after homologous transmission, for instance, in studies using the oral route of infection and low dosage inoculum [79,83,84]. 3. Fish prion proteins 3.1. Duplicated PrPs in bony fish PrP genes from land vertebrates have typically been cloned using conventional procedures such as cDNA hybridization and PCR screening. In contrast, fish PrP genes could not be isolated with these methods due to their large DNA sequence divergence. The first fish PrP homologue was identified in Fugu rubripes by searching genome databases for ORFs encoding key structural features of PrPs [64]. Subsequent cloning and sequencing of PrP genes in salmon, trout, carp, stickleback, sea bream and zebrafish [67,68,85,86] revealed that bony fish possess two PrP orthologs, PrP-1 and PrP-2, which probably arose during a fish-specific genome duplication [87]. Northern and Western Blot analyses indicated that the fish proteins are expressed at particularly high levels in adult fish brains, as well as in muscle, skin, heart and gills [67]. Although quite variable in size and sequence, PrP1 and PrP-2 present characteristic features of mammalian PrPs: an Nterminal signal peptide, conserved N-glycosylation sites, two cysteine residues that form a disulphide bond, a central hydrophobic stretch, and the C-terminal signal for the attachment of a GPI-anchor (Fig. 1). An intriguing feature of fish PrPs is the presence of a highly conserved motif of 13 amino acids in length between the repetitive region and the hydrophobic stretch [67]; the structural or functional importance of this motif is not evident from its sequence. The repetitive domains of fish PrPs are easily recognizable but at the same time clearly different from those of tetrapod PrPs. For example, while a constant number of nearly identical octarepeats is typical for mammals (or hexarepeats in birds), in fish the number and length of degenerate repeats vary from species to species. Nevertheless, all PrP repeats can be considered variations of two basic types of repeats (A and B), which evolved differentially in every vertebrate class [67]. The sequence of the globular domain also has diverged considerably between fish and mammals, owing to high rates of amino acid 406 E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 substitutions, insertions and deletions. Notably, these changes predominantly affect residues outside key structural motifs, and are not expected to affect the globular fold [67]. Apart from PrP-1 and PrP-2, other genes partially resembling PrP have been reported in fish. Originally described as PrP-like and Shadoo [88,89], they are also known as PrP-1-rel and PrP-2-rel because of their genomic location, directly downstream of PrP-1 and PrP-2, respectively [67]. The encoded products are GPI-anchored polypeptides of only 150–180 amino acids in length, expressed in brain tissues, and containing the characteristic PrP hydrophobic stretch. Further similarities to PrPs include a β-strand motif in PrP-rel-1 identical to the first β-strand of PrP, and the presence of degenerated B- and Alike repeats in PrP-rel-2. This situation is reminiscent to that of Prnp and its genomic neighbor Doppel. Thus, PrP-rels, like Doppel, may have arisen from the tandem duplication of an ancestral PrP gene, and then diverged as a result of high substitution rates and the differential loss of motifs. Interestingly, although PrP-rels are predicted to be unstructured proteins, their mammalian homologue Shadoo has recently received attention because it appears to influence biological and pathogenic activities of PrP in vivo [90,91]. 3.2. Zebrafish PrP-1 and PrP-2 The characterization of PrP homologues in the zebrafish opened the exciting possibility to study prion biology in a new model organism. Like other bony fish, the zebrafish has two prion protein orthologs, PrP-1 and PrP-2, mapped to chromosomes 10 and 25, respectively [67]. Genomic studies revealed that the two genes originated from the duplication of a large chromosomal block that corresponds to those containing the Prnp gene in mammals. That both gene copies are fully functional has been biochemically confirmed in fish and mammalian cells [92,93]. In these studies, it was shown that the encoded polypeptides become properly glycosylated and attached to the plasma membrane via a GPI-anchor. As is often the case with duplicated genes [94], PrP-1 and PrP-2 are thought to have retained complementary subsets of ancestral PrP functions. Accordingly, zebrafish PrPs have evolved distinct spatiotemporal patterns of embryonic expression: while PrP-1 is ubiquitously expressed at high levels in very early stages of embryogenesis (blastula and gastrula stages), PrP-2 is strongly upregulated at later stages in developing neurons [92]. At later larval stages and into adulthood, both genes continue to be expressed in the CNS, although PrP-2 at much higher levels than PrP-1. The differences between the expression patterns of PrP-1 and PrP-2 are due to the divergent evolution of DNA transcriptional regulatory sequences upstream of each gene (Barth and Málaga-Trillo, unpublished data). Interestingly, the expression of zebrafish PrPs (particularly PrP-2) in the developing nervous system bears striking similarities to that of PrP in the mouse embryo [32]. It remains to be determined if, like PrP-1, mammalian PrP plays a role during blastula and early gastrula stages. 3.3. PrP loss of function phenotypes in zebrafish embryos One of the many experimental advantages of the zebrafish is the possibility to knockdown the expression of any given gene using morpholino antisense oligonucleotides. When microinjected into early fish embryos, morpholinos rapidly bind to their target mRNAs and block their translation [95]. Recently, this method was successfully applied to study PrP loss of function in the zebrafish. In contrast to the absence of clear phenotypes in PrP knockout mice, knockdown of zebrafish PrPs revealed that these genes play essential roles during distinct phases of embryonic development. Specifically, PrP-1 knockdown causes lethal embryonic arrest during gastrulation, whereas PrP-2 knockdown severely impairs brain development. Notably, the PrP-1 phenotype can be rescued by PrP-1 or PrP-2 overexpression, indicating that although the two proteins are expressed in different developmental contexts, they share essentially the same biological activity. Furthermore, the ability of mouse PrP to partially revert the PrP-1 phenotype strongly suggests that this activity of PrP is conserved between fish and mammals [92]. It is therefore surprising that the loss of such an important function would not be evident in PrP knockout mice, especially since PrP is expressed at high levels in the mouse embryo. A plausible explanation for this discrepancy is that the mouse knockout phenotype may become masked by genetic compensation or developmental plasticity [96]. The relative simplicity of the zebrafish gastrula makes the PrP-1 phenotype an excellent choice to analyze conserved mechanisms of PrP function at the molecular and cellular levels. Detailed characterization of PrP-1 knockdown embryos revealed that the developmental arrest was caused by the loss of tissue cohesion and the consequent failure to carry out epiboly, an early morphogenetic cell movement. Because epiboly relies on the control of cell–cell adhesion, this finding led to the discovery that PrP-1 regulates the stability of E-cadherin/ßcatenin adhesive complexes at the plasma membrane [92]. Ongoing biochemical work indicates that the mechanism behind this effect is the ability of PrP-1 to control the levels and membrane localization of ß-catenin as well as the levels and phosphorylation state of Srcrelated kinases (Sempou and Málaga-Trillo, in preparation). Related unpublished in vivo pharmacological data suggest that PrP-1 acts to prevent the degradation of some of these proteins. These results are in agreement with long proposed roles of PrP as a signal transduction molecule [46]. Concretely, the experiments with zebrafish gastrulae indicate that a complex signaling cascade is initiated by the interaction between PrP-1 molecules across opposite cell membranes. Accordingly, in vitro experiments show that fish, amphibian, avian and mammalian PrPs share the intrinsic ability to engage in such homophilic trans-interactions, promoting the formation of cell–cell contacts, and triggering intracellular signaling via Srcrelated kinases [92]. Of special interest within the scope of this review is the finding of heterologous interactions between zebrafish and mammalian PrPs [92], which raise the question of whether PrP conversion and TSE transmission can take place across such distant species. Altogether, the combined work in zebrafish embryos and mammalian cells indicate that PrPs play crucial roles in cell–cell communication (Fig. 1). These data are in line with the notion that PrP provides protective signals required for neuronal growth and/or survival [97,98]. Understanding the precise nature of such signals will be crucial to uncover the early cellular mechanisms of prion pathogenesis in the mammalian brain. 4. Risk of prion disease in fish 4.1. TSEs and the food chain The appearance of BSE in Europe has been causally linked to the common practice of feeding cattle with meat and bone meal (MBM) of bovine origin [99]. Approximately 200,000 BSE cases were reported between 1985 and 2000, and it has been estimated that roughly three million infected animals could have entered the human food chain before developing a clinical disease [81]. Due to its low cost and high nutritional value, MBM has become an important ingredient in the diet of non-ruminant livestock and farmed fish. Hence, there is the theoretical risk that farmed animals unintentionally fed with prioncontaminated MBM might develop TSEs or serve as asymptomatic carriers of a disease [100]. Since BSE was initially reported, the European Commission (EC) has implemented a comprehensive set of TSE risk-reducing measures to protect humans from BSE, as well as to control and eventually eradicate prion diseases in animals. Among these, a total EU-wide feed ban on the use of rendered animal proteins in the nutrition of all animals was imposed, with minor exceptions (i.e. use of fish feeds in the diets of non-ruminants like pigs and chicken). E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 Experimental demonstration that TSEs can propagate in a new host without typical pathognomonic brain lesions, clinical disease or fatality, has led to the conclusion that prion-infected individuals could go undetected within an exposed population [100]. This implies that farmed animals may carry the disease but not develop it until they are slaughtered. Human consumption of meat from these animals would, in turn, give rise to repeated infection cycles long before clinical signs can be recognized. Moreover, the relatively recent discovery of unconventional TSE phenotypes, such as atypical scrapie in sheep and goats [101,102], and bovine amyloidotic spongiform encephalopathy (BASE) in cattle [103], usually displaying decreased or altered PrPSc resistance to proteolytic treatment and less pronounced clinical signs, has further made it difficult to define prion diseases in absolute terms and thus, to implement effective rules to eliminate them. 4.2. Risk of prion transmission to fish According to the aforementioned arguments, fish, poultry and pigs represent potential candidate hosts for prion infection. In order to assess the risk of prion transmission to fish, five main parameters should be considered: 1. the potential use of prion-contaminated MBM as a nutritional supplement in aquaculture, 2. the consumption of infected farmed fish by humans, 3. the use of feeds or other products (i.e. gelatin, milk replacers) derived from TSE-affected fish for mammalian or piscine nutrition, 4. the use of fishmeals cross-contaminated with MBM in mammalian diets, and 5. the escape of infected fish, or the release of infected waste from aquaculture facilities into the marine environment. These scenarios are depicted in Fig. 2. With fish mariculture turning into a very important and rich protein source for humans, consumers may become concerned about the possibility of farmed fish developing prion disease, or serving as passive carriers of prion infectivity. All farmed fish receive commercial feeds containing 40–55% protein, and since some of it is likely to be of animal origin, the possibility of feed contamination with mammalian prions cannot be excluded. Interestingly, in the late 1980s, total fish feed production in the UK was in the order of 75,000 tons per year, with the use of about 3750 tons of MBM in fish feed [104]. However, how realistic is the thesis that fish prion proteins may misfold and form abnormal aggregates? The striking conservation of structural motifs between mammalian and piscine PrP molecules 407 makes it rather difficult to exclude such a possibility. Regarding the convertibility potential of fish PrPs, it has been suggested that the extended type A-repeat domain in sea bream PrP-1 and other fish PrPs could confer on these proteins the ability to self-polymerize and aggregate [85]. In addition, it has been noted that motifs similar to those observed in fish PrPs can also be identified in unrelated proteins such as the piscine annexin 11 and the mammalian lectin L-29 [77]. Notably, both these proteins display self-aggregation properties, and annexin 11 is, like PrP, a neuro-specific membrane protein. 4.3. Transmission studies in commercially important fish species At present, only a few studies have explored the experimental transmission of TSEs to teleost fish. The first attempts in this direction involved the oral and parenteral inoculation of two commercially important fish species, rainbow trout and turbot, with the 139A mouse-adapted scrapie strain [105]. Although the fish displayed no clinico-histopathological signs during the three month-experimental period, a mouse bioassay revealed that they carried residual infectivity. Specifically, mice intracranially inoculated with trout intestinal extracts one day after oral challenge were positive for brain PrPSc deposition approximately 200 days post inoculation (p.i.), despite the absence of clinical symptoms. Similar results were obtained when mice were intracerebrally (i.c.) inoculated with spleen extracts from trouts and turbots, 15 days after parenteral inoculation of the fish with scrapie material. Finally, brain tissue from parenterally inoculated turbots 15 and 90 days after challenge was also able to elicit PrPSc accumulation in the brains of recipient mice, without causing clinical disease. Recently, we reported data on the oral transmission of BSE and scrapie to gilthead sea bream (Sparus aurata) [106]. Interestingly, at two years p.i., a number of fish that had been force-fed BSE- or scrapie-infected brain homogenates developed abnormal plaque-like deposits in their brains. Specifically, the brains of two out of five fish inoculated with scrapie developed signs of abnormal protein aggregation at 24 months p.i. These aggregates were positively stained with polyclonal antibodies raised against fish PrPs, but showed no proteinase K (PK)-resistance or Congo red birefringence (Fig. 3A). The brains of the BSE-challenged fish, however, displayed a much more striking picture, having already developed the first signs of abnormal deposition at eight months p.i. A general progression in size, PK-resistance and morphological features was observed thereafter, resulting in an impressive number of aggregates in all the brain regions examined at 24 months p.i. Three out of five fish sacrificed at this time point showed 500–800 deposits per brain section each, 70– 85% of which were PK-resistant and had a mean diameter of 30 μm. These aggregates were PAS-positive, congophilic and birefringent in polarized light, indicating an amyloid or amyloid-like fibrillar structure (Fig. 3B). In contrast to the TSE-challenged individuals, no signs of abnormal aggregation or any other lesions were observed in the brains of the control fish “challenged” with bovine or ovine brain homogenates prepared from healthy animals. Altogether, the development of abnormal brain deposits in BSE-challenged sea bream constitutes an unprecedented histopathology in fish. 4.4. Neuropathology in TSE-challenged sea bream Fig. 2. Prion diseases and the food chain. Potential scenarios for the transmission of prions between mammals and fish are illustrated. Arrows indicate the direction of possible transmission events. The lesions observed in the brains of the BSE-infected sea bream share some remarkable similarities with those typically seen in prionaffected mammalian brains. For instance, the deposits in fish brains were only detectable where neuronal parenchyma was present, a feature greatly resembling the localization of prion amyloid plaques in mammals [107]. The most extensively affected regions included the cerebellum (which displayed a mammal-like deposition motif [108]), the lateral nucleus of the ventral telencephalic area (a fish counterpart to the basal nucleus of Meynert in mammals [109]), the lateral 408 E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 Fig. 3. Experimental transmission of prions between mammals and fish. Inoculation of fish with ovine (A), bovine (B) and mouse (C) prions and their various outcomes. The photographs on the upper right corner illustrate the development of abnormal amyloid deposition in the optic tectum of BSE-challenged sea bream, using the periodic acid Schiff (PAS) staining reaction. The right image (scale bar = 10 μm) is a zoom-in of the lesion indicated in the left image (scale bar = 100 μm). i.c., intracerebrally. telencephalic pallium (homologue to the mammalian hippocampus [110]), and finally the thalamus and diencephalon. Notably, no spongiosis was observed in any of the brain regions examined. However, cases of TSE subtypes that develop little or no vacuolation have also been reported for human prion diseases [111]. Interestingly, the neurodegeneration observed in BSE-challenged sea bream affected only neurites, as opposed to the “classically defined” degeneration of neuronal somata. Based on the morphological progression of the abnormal deposits and the level of neurite involvement, aggregates were classified into two main categories. “Pre-mature” deposits, observed mostly at earlier time points and consisting mainly of dystrophic neurites that have lost their coherence, appear to mark initial stages of neurodegeneration. “Mature” deposits, observed 24 months p.i., seem to result from the complete deconstruction of fibers, leading to the development of a more homogenous and flocculated extracellular material, with increased PrP-immunoreactivity, congophilia and birefringence in polarized light. The two types of aggregates in the brains of BSE-challenged sea bream could be interpreted as different stages of pathogenesis, with the former temporally preceding the latter. For scrapie-challenged sea bream, the distribution and morphology of the observed neuropathology could not be evaluated and classified as mentioned earlier, due to the limited number of fish developing abnormal deposits in this group. 4.5. Absence of “clinical symptoms” in fish: presymptomatic carrier state? Visual examination of challenged fish throughout the post inoculation period revealed no signs of altered swimming behavior or eating habits. However, previous studies have shown that species once thought to be resistant to certain TSE strains can be life-long carriers of the infection without ever developing a clinical disease. For example, the absence of clinical symptoms has been reported in mice following the primary transmission of sheep and hamster scrapie. However, sequential passage of infectious material did result in the development of clinical disease [82]. Interestingly, differential susceptibility to specific pathogens is also known among piscine species. For instance, closely related fish species may exhibit differential resistance to viral infections, as illustrated by the development of clinical disease in sea bass but not in sea bream infected with nodavirus [112,113]. On the other hand, the absence of clinical symptoms in BSE-challenged sea bream could be related to the remarkable ability of teleost fish to continuously produce new neurons throughout their lifetime. It is well known that adult fish can regenerate destroyed retinal tissue, optic nerve axons, and degenerated brain stem neurites, resulting in functional recovery of these structures [114,115]. Given the fact that farmed fish remain in aquaculture facilities up to 20 months before reaching the required commercial weight, the lack of clinical disease in infected animals may become a crucial issue of unanticipated relevance for public health. 4.6. Differential susceptibility of fish species to prion infection In a parallel study, we inoculated European sea bass (Dicentrarchus labrax) with the same infectious material, namely BSE and scrapie. Interestingly, in contrast to the sea bream, none of the challenged individuals showed any clinical symptoms or histopathological evidence of abnormal deposition at any point throughout the time course of the study. Sea bass and sea bream are both Perciformes, the largest class of teleosts. Although they belong to different families, Sparidae (sea bream) and Moronidae (sea bass), the two species are E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 409 quite closely related at the phylogenetic and anatomical levels. Since the amino acid sequence of sea bass PrP is not available, it is not clear how it might differ from that of sea bream PrP. However, the two proteins display different reactivities against anti-fish PrP antibodies. Conceivably, such differences could confer on the sea bass PrP resistance to the PrPSc in the inoculum, and thus, reinforce the “transmission barrier” in these experiments. Another factor possibly involved in the different reaction of the two species to prion inoculation might be the ease with which exogenous PrPSc can cross the intestinal barrier and reach the CNS [116]. In teleosts, like in other vertebrates, oral pathogens are initially confronted by activated immune cells on the enteric wall. Macrophages, whose efficiency contributes to the resistance against pathogens, play an important role in this process. While in sea bass, enteric phagocytes seem to become extensively activated upon microbial attack [117], it has been reported that the intestinal mucosa of the sea bream may be relatively permissive to infection by certain bacteria [118]. Although plausible, these hypotheses do not entirely explain the differential sensitivity of the two species to prion infection, since cases of pathogenesis in which the clinically affected species is the sea bass and not the sea bream are also known [112]. It is noteworthy, however, that the different susceptibility of the two species can be regarded as an internal control in our transmission experiments. Specifically, the fact that despite the close relationship between the two species, only sea bream were affected, strengthens the hypothesis that the observed neuropathology represents a novel piscine amyloidosis triggered by mammalian prions. Thus, transmissibility of prions to fish exhibits species specificity, a common feature of TSEs. the GI tract may serve as a potential portal for exogenous PrPSc, due to the high PrP mRNA levels detected in the posterior intestine of zebrafish larvae [68]. In mammals, studies have shown that the pathogenic agent may invade the CNS directly, by exploiting the extended autonomous intestinal innervation [120]. In particular, it has been proposed that BSE neuroinvasion may take place via the parasympathetic nerve fibers of the vagus nerve into the brain stem of orally-challenged cattle [81]. Such a route of dispersion requires the adequate expression of endogenous PrPC within the peripheral nervous system [121]. In the sea bream, immunohistochemical examination of intestinal tissue revealed three main regions of intense PrP-immunopositivity, namely the serosa, the myenteric plexus (corresponding to Auerbach's plexus in mammals) and the submucous plexus (a homologue to mammalian Meissner's plexus). Both the myenteric and the submucous plexus contain many parasympathetic nerves and ganglia, one of which is a nucleus of the vagus nerve running to the medulla oblongata in the brain stem of mammals and teleosts. The hypothesis that the infectious agent might use the enteric nervous system to invade the fish CNS is supported by the progressive changes observed in the regional distribution of abnormal deposits in sea bream brain [81]. Initially, the region affected is the posterior part of the brain, with the brain stem displaying the most severe lesion profile. Subsequently, anterior anatomical regions show signs of abnormal aggregation as well, although posterior brain areas continue to be affected. Interestingly, it has been proposed that direct neuroinvasion of the infectious agent following per os challenge might be related to the absence of clinical symptoms and the reduced attack rate of the transmitted disease [122], an observation that is in agreement with the asymptomatic amyloidosis developed by BSE-challenged sea bream. 4.7. Possible routes of neuroinvasion: tissues harboring infectivity 5. Future directions Immunohistochemical examination of peripheral tissues from TSEinoculated sea bream including spleen and intestine revealed no signs of abnormal protein accumulation at any of the time points tested. Therefore, it is likely that the transit of prions from the fish gastrointestinal (GI) tract to the CNS might have occurred long before the first time point of sacrifice, perhaps only a few hours or days after oral challenge, leaving no trace of infectivity in the peripheral tissues tested thereafter. Transmission studies in rainbow trout and turbot have shown that specific binding of exogenous PrPSc (139A scrapie strain) can be detected in the intestinal mucosa for up to seven days p.i., despite the absence of active uptake by the epithelial cells [105,119]. Preliminary data from a more recent study demonstrate the importance of residual infectivity, suggesting that after oral challenge with the ME7 mouse-adapted scrapie, prions can cross the intestinal wall of the sea bream and be delivered from the peritoneal cavity into the spleen (Figueras et al., unpublished data). More specifically, inoculated fish were sacrificed at regular time points, and various tissues, including spleen and intestine, were used to intracranially inoculate mice (Fig. 3C). The TSE-challenged fish never showed histopathological or clinical symptoms throughout the 18 monthexperimental period. However, serial passage to mice of tissues from fish sacrificed one week p.i. induced vacuolation in the brains of some of the mice, 300 days after the intracranial inoculation. Concretely, spongiosis without clinical disease was detected in five out of 19 recipient mice inoculated with scrapie-infected fish spleens, and in 12 out of 17 individuals challenged with scrapie-infected fish intestines. These findings suggest, on one hand, that the TSE agent might have been able to reach the fish brain in a remarkably short period of time, either after peripheral propagation or even by direct neuroinvasion. On the other hand, this study emphasizes the need for a thorough examination of peripheral tissues in asymptomatic individuals following prion inoculation. There is no doubt that the intestine represents the main entry point for oral pathogens. In zebrafish, for example, it has been proposed that 5.1. Monitoring for infectious neurodegenerative disease in farmed fish The rapid worldwide expansion of fish farming calls for the implementation of efficient measures to ensure animal health and food safety. The fact that pathogens can spread considerably more easily through flowing water than between land animals is one of the main difficulties when dealing with aquatic diseases. Over the last years, the aquaculture industry has been faced with an increasing number of infectious diseases affecting both farmed and feral populations. Surveillance and monitoring programs have therefore been developed in many countries, in an attempt to prevent and control the emergence of new fish pathologies [123]. The list of notifiable diseases varies depending on the specific needs of each region. In Europe, all countries are asked to comply with the Council Directive (2006/88/EC) on health requirements for aquaculture animals. Nevertheless, it would be sensible to enrich the list of large-scale spot checks in fish farms by including the histological examination of fish brains, in order to detect possible signs of feed-induced or even spontaneous PrPSc amyloidosis. As described earlier, BSE-infected sea bream developed the first signs of abnormal aggregation already at eight months p.i., a time period that lies within the grow-out phase of fish breeding. Moreover, the fact that these animals remained asymptomatic throughout the experimental period raises the possibility that affected fish may generally go undetected before ending up on our plates. 5.2. Further transmission studies in commercially important fish species Gilthead sea bream and European sea bass are the two most important farmed fish in the Mediterranean Sea, with the main producers being Greece, Turkey, Spain, France and Italy. In 2004, the annual production levels of both species reached 160,000 tons worldwide. Therefore, assessing the risk of TSE transmission to these fish is an issue of major importance for public health. On a global scale, the practice of aquaculture is experiencing steady growth rates, driven 410 E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 by the increasing demand for aquatic food products. According to projections by the Food and Agriculture Organization of the United Nations (FAO), the global production of farmed fish will have to reach 80 million tons by 2050, in order to maintain the current level of fish consumption per capita. Amazingly, not all countries have introduced laws to regulate the use of rendered mammalian proteins for fish nutrition and, in some cases, those that have, have incurred in minor and major violations [124]. Hence, the likelihood that contaminated MBMs might have been inadvertently fed to farmed fish is considerably high. In this regard, the amyloidosis observed in the brains of BSE-fed sea bream demonstrates that certain fish species can develop neuropathology following ingestion of prion-contaminated meals. Therefore, further transmission studies with other commercially important fish species, including salmon (Salmo salar), trout (Oncorhynchus mykiss), and carp (Cyprinus carpio) should be conducted in order to evaluate the likelihood of acquiring TSEs upon consumption of infected feeds (Fig. 4). In addition, primary TSE transmission studies in piscine species that are considered future candidates for mariculture would be advisable, especially those closely related to sea bream, such as the common dentex (Dentex dentex, Sparidae) and the red porgy (Pagrus pagrus, Sparidae). 5.3. Zebrafish as a model for prion pathobiology Clearly, transmission experiments in commercially important species are necessary to directly assess the risk that TSE-infected fish would pose to humans. On the other hand, addressing the cellular and molecular aspects of prion transmission to fish requires the use of established model organisms. The zebrafish is best fitted for this purpose, not only because of its well-characterized neuroanatomy and neural development, but also because of the availability of useful research tools, such as mutant strains, GFP-transgenic lines, specific antibodies, cell lines and a fully sequenced genome. Additionally, zebrafish larvae are optimally suited for the study of neurological diseases because they display quantifiable neuropathological and behavioral phenotypes similar to those observed in humans [125]. It would be particularly interesting to know whether scrapie or CJD prions can be efficiently inoculated to transgenic zebrafish expressing mouse or human PrPs. Similarly, expression of pathogenic PrP mutants in wildtype and transgenic fish could be of great value in elucidating the genetic basis of prion disease. Engineering of these animals, currently underway, will provide an ideal tool to systematically analyze prion replication in fish brains, as well as the development of PrP aggregates, brain lesions, and possibly clinical disease. Another fascinating perspective is the possibility to study the effect of exogenously added prions on zebrafish embryos. Such experiments could significantly facilitate the analysis of PrPC and PrPSc behavior in vivo, thus contributing valuable insights into the relationship between prion conversion and PrP dysfunction. Altogether, transmission experiments in zebrafish could help to effectively evaluate the susceptibility of fish to TSEs, as well as the mechanistic link between PrP function and prion pathogenesis. Fig. 4. Future scenarios for transmission studies. Examples of primary and secondary inoculation studies required for the evaluation of prion transmissibility between piscine and mammalian species. Tg = transgenic, wt = wildtype, ZF = zebrafish, boPrP = bovine PrP, ovPrP = ovine PrP, huPrP = human PrP, moPrP = mouse PrP. E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 411 5.4. Assessing transmissibility and convertibility 5.5. Dissecting the physiological roles of PrP in the zebrafish Issues of prion transmissibility between fish and mammals need to be given full consideration because of the possibility that fish fed with prion-contaminated MBM may further transmit TSEs through the food chain. To evaluate this scenario, secondary transmission experiments between BSE-affected fish brains and mice carrying the bovine Prnp gene are currently underway. Further studies in transgenic fish will also help assess the risk of potential transmission of piscine amyloidoses to mammals or to other fish species (Fig. 4). For instance, the use of transgenic zebrafish carrying the bovine Prnp, or mice carrying the gene coding for fish PrP on a PrP knockout background would answer a plethora of questions concerning not only host-related pathogenic determinants but also the specific conditions required for certain intra-species transmissions. The discovery of residual prion infectivity in fish intestines and spleens is also a matter of great concern, as it suggests that even fish species resistant to prion infection might harbor infectivity shortly after ingestion of contaminated meals, thus being able to transmit the disease to consumers. Similar studies should be undertaken employing a variety of fish species and TSE strains, including atypical scrapie and BSE, to assess residual infectivity levels not only in spleen and intestine, but also in blood and other “high risk” tissues (Fig. 4). Furthermore, bioassays should also include “humanized” mice in order to directly assess the risk of transmission to humans. Another basic question that requires further clarification is the convertibility potential of fish PrPs. Ongoing experiments using the sensitive protein misfolding cyclic amplification (PMCA) method [126,127] may reveal whether piscine PrPs have the intrinsic ability to misfold and aggregate. In addition, the use of TSE-infected cell lines could become an important tool to investigate possible molecular interactions between mammalian PrPSc and fish PrPC, leading to the conversion of piscine PrPs. Aside from transmissibility and convertibility issues, the composition of the deposits found in the brains of the BSE-challenged sea bream deserves further examination. Concretely, the identification of other aggregated molecules in brain deposits could help define the properties of these novel amyloid plaques, and perhaps their mechanistic connection to neurodegeneration. For instance, previous data on the composition of amyloid deposits present in degenerating neuronal tissue have suggested that the components of such aggregates may represent the outcome of the degeneration process rather than the cause of neuronal cell death [128]. Thus, the study of novel proteinaceous deposits may contribute new leads to the debate about the causal link between neurotoxicity and protein oligomerization. The molecular basis of prion disease is closely associated with the structural and functional properties of PrP. Therefore, deep understanding of prion pathogenesis in fish and mammals cannot be reached without basic knowledge about the physiological roles of PrP. The knockdown phenotypes described in zebrafish constitute the first experimental evidence that the lack of PrP can cause dramatic abnormalities in a whole organism [92]. In addition, the finding that PrPs regulate cell–cell communication via known cellular pathways provides an ideal opportunity to establish how PrP may exert its proposed neuroprotective role (Fig. 5). It is noteworthy that some of the molecules affected in PrP-1 knockdown embryos (Src-kinases, Ecadherin, β-catenin, etc.) are not only essential for early zebrafish development but also play important roles in neuronal function and during neurodegeneration [96]. Therefore, it will be crucial to further characterize the signaling pathways under the control of PrP-1 using biochemical, cell biological, proteomic and imaging tools. Such experiments are currently being carried out in fish embryos and embryonic stem cells. Ideally, this basic knowledge should be integrated into the design and interpretation of studies addressing the function of PrP in mammalian cells, as well the experimental transmission of prion disease between fish and mammals. Characterization of the PrP-2 knockdown phenotype is currently in progress. Preliminary results show that PrP-2 is required for the proliferation and differentiation of different groups of embryonic neurons (Málaga-Trillo and Luncz, unpublished data). A similar phenotype has been reported for PrP knockout mice [41], although the defect does not impair neural development to the extent seen in the zebrafish [92]. Because PrP-2 also shares the ability to elicit cell– cell adhesion and intracellular signaling, it is likely to regulate neural development via the same molecular pathways controlled by PrP-1 in the early gastrula. Thus, morphological and molecular analyses of the PrP-2 phenotype may provide valuable mechanistic insights into the roles of PrP during neuronal differentiation, axon growth, and synaptogenesis. Integrating the emerging zebrafish data within the vast amount of information available from mammalian cells and knockout mice promises to become a lengthy but rewarding task. 6. Concluding remarks For many years, prion researchers have relied heavily on the use of mammalian models as experimental systems. This strategy, and particularly the development of genetically modified mice, has led to outstanding discoveries in the field. At the same time, the Fig. 5. Exploring the roles of PrP in health and disease. It has been proposed that neuronal death results from failure by PrP to provide an unknown neuroprotective function. It remains to be clarified how the various cellular and molecular processes influenced by PrP in vivo may converge into such a neuroprotective role. 412 E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 experimental difficulties associated with the study of TSEs have exposed the need for new animal models in prion biology. Recently, researchers have explored the utility of fish to address questions concerning the physiological roles of PrP and the risk of TSEs in these organisms. An interesting picture has emerged, suggesting that transmission of prions could potentially take place between commercially important fish and mammals, although the implications of these findings for human and animal health remain to be clarified. Much remains to be done in order to confirm or rule out the existence of bona fide prions in fish. Nevertheless, the use of fish models in TSE research will likely uncover novel aspects of prion transmission across a wide range of hosts, relevant to human health. In addition, the finding that fish and mammalian PrPs share important roles in cell– cell communication might help elucidate the molecular mechanisms by which prions induce neuronal death, setting the stage for the identification of new anti-prion therapeutic targets. Acknowledgements This work is supported by a DFG grant (MA 2525/2-1) and University of Konstanz funds awarded to EMT. ES, TS and AF are funded by EU grants (QLK5-2002-00866 for ES, TS, AF and FOOD-CT-2004-506579 for ES, TS). ES is a scholar of the Greek States Scholarships Foundation (IKY), and the sea bream and sea bass transmission experiments were part of her doctoral thesis. AF would like to thank Raquel Aranguren and Beatriz Novoa for contributing to the secondary transmission study of ME7infected sea bream tissues to mice. References [1] S.B. Prusiner, The prion diseases, Brain Pathol. 8 (1998) 499–513. [2] J. Collinge, Prion diseases of humans and animals: their causes and molecular basis, Annu. Rev. Neurosci. 24 (2001) 519–550. [3] C.M. Poser, Notes on the history of the prion diseases, Part I Clin. Neurol. Neurosurg. 104 (2002) 1–9. [4] A. Aguzzi, F. Baumann, J. Bremer, The prion's elusive reason for being, Annu. Rev. Neurosci. 31 (2008) 439–477. [5] C.M. Poser, Notes on the history of the prion diseases, Part II Clin. Neurol. Neurosurg. 104 (2002) 77–86. [6] W.J. Hadlow, Scrapie and kuru, Lancet 2 (1959) 289–290. [7] I. Klatzo, D.C. Gajdusek, V. Zigas, Pathology of kuru, Lab. Invest. 8 (1959) 799–847. [8] J. Cuillé, P.L. Chelle, La maladie dite tremblante du mouton est-elle inoculable? CR Acad. Sci. 203 (1936) 1552–1554. [9] D.C. Gajdusek, C.J. Gibbs, M. Alpers, Experimental transmission of a kuru-like syndrome to chimpanzees, Nature 209 (1966) 794–796. [10] C.J. Gibbs Jr., D.C. Gajdusek, D.M. Asher, M.P. Alpers, E. Beck, P.M. Daniel, W.B. Matthews, Creutzfeldt–Jakob disease (spongiform encephalopathy): transmission to the chimpanzee, Science 161 (1968) 388–389. [11] H. Diringer, Transmissible spongiform encephalopathies (TSE) virus-induced amyloidoses of the central nervous system (CNS), Eur. J. Epidemiol. 7 (1991) 562–566. [12] B. Chesebro, Introduction to the transmissible spongiform encephalopathies or prion diseases, Br. Med. Bull. 66 (2003) 1–20. [13] I.H. Pattison, Resistance of the scrapie agent to formalin, J. Comp. Pathol. 75 (1965) 159–164. [14] S.B. Prusiner, Novel proteinaceous infectious particles cause scrapie, Science 216 (1982) 136–144. [15] T. Alper, W.A. Cramp, D.A. Haig, M.C. Clarke, Does the agent of scrapie replicate without nucleic acid? Nature 214 (1967) 764–766. [16] J.S. Griffith, Self-replication and scrapie, Nature 215 (1967) 1043–1044. [17] B. Chesebro, B. Caughey, Scrapie agent replication without the prion protein? Curr. Biol. 3 (1993) 696–698. [18] C.I. Lasmezas, J.P. Deslys, O. Robain, A. Jaegly, V. Beringue, J.M. Peyrin, J.G. Fournier, J.J. Hauw, J. Rossier, D. Dormont, Transmission of the BSE agent to mice in the absence of detectable abnormal prion protein, Science 275 (1997) 402–405. [19] L. Manuelidis, Z.X. Yu, N. Barquero, B. Mullins, Cells infected with scrapie and Creutzfeldt–Jakob disease agents produce intracellular 25-nm virus-like particles, Proc. Natl Acad. Sci. USA 104 (2007) 1965–1970. [20] M.E. Keyes, The prion challenge to the ‘central dogma’ of molecular biology, 1965–1991, Stud. Hist. Philos. Biol. Biomed. Sci. 30 (1999) 1–19. [21] C. Weissmann, The state of the prion, Nat. Rev. Microbiol. 2 (2004) 861–871. [22] S.B. Prusiner, Molecular biology of prion diseases, Science 252 (1991) 1515–1522. [23] J.T. Jarrett, P.T. Lansbury Jr., Seeding “one-dimensional crystallization” of amyloid: a pathogenic mechanism in Alzheimer's disease and scrapie? Cell 73 (1993) 1055–1058. [24] D.C. Bolton, M.P. McKinley, S.B. Prusiner, Identification of a protein that purifies with the scrapie prion, Science 218 (1982) 1309–1311. [25] B. Oesch, D. Westaway, M. Walchli, M.P. McKinley, S.B. Kent, R. Aebersold, R.A. Barry, P. Tempst, D.B. Teplow, L.E. Hood, et al., A cellular gene encodes scrapie PrP 27–30 protein, Cell 40 (1985) 735–746. [26] B. Chesebro, R. Race, K. Wehrly, J. Nishio, M. Bloom, D. Lechner, S. Bergstrom, K. Robbins, L. Mayer, J.M. Keith, et al., Identification of scrapie prion protein-specific mRNA in scrapie-infected and uninfected brain, Nature 315 (1985) 331–333. [27] D.R. Borchelt, M. Scott, A. Taraboulos, N. Stahl, S.B. Prusiner, Scrapie and cellular prion proteins differ in their kinetics of synthesis and topology in cultured cells, J. Cell Biol. 110 (1990) 743–752. [28] K.M. Pan, M. Baldwin, J. Nguyen, M. Gasset, A. Serban, D. Groth, I. Mehlhorn, Z. Huang, R.J. Fletterick, F.E. Cohen, et al., Conversion of alpha-helices into betasheets features in the formation of the scrapie prion proteins, Proc. Natl Acad. Sci. USA 90 (1993) 10962–10966. [29] K. Basler, B. Oesch, M. Scott, D. Westaway, M. Walchli, D.F. Groth, M.P. McKinley, S.B. Prusiner, C. Weissmann, Scrapie and cellular PrP isoforms are encoded by the same chromosomal gene, Cell 46 (1986) 417–428. [30] R. Riek, S. Hornemann, G. Wider, R. Glockshuber, K. Wuthrich, NMR characterization of the full-length recombinant murine prion protein, mPrP (23–231), FEBS Lett. 413 (1997) 282–288. [31] K.J. Knaus, M. Morillas, W. Swietnicki, M. Malone, W.K. Surewicz, V.C. Yee, Crystal structure of the human prion protein reveals a mechanism for oligomerization, Nat. Struct. Biol. 8 (2001) 770–774. [32] J. Manson, J.D. West, V. Thomson, P. McBride, M.H. Kaufman, J. Hope, The prion protein gene: a role in mouse embryogenesis? Development 115 (1992) 117–122. [33] M.J. Ford, L.J. Burton, R.J. Morris, S.M. Hall, Selective expression of prion protein in peripheral tissues of the adult mouse, Neuroscience 113 (2002) 177–192. [34] W. Goldmann, G. O'Neill, F. Cheung, F. Charleson, P. Ford, N. Hunter, PrP (prion) gene expression in sheep may be modulated by alternative polyadenylation of its messenger RNA, J. Gen. Virol. 80 (Pt 8) (1999) 2275–2283. [35] D.A. Harris, P. Lele, W.D. Snider, Localization of the mRNA for a chicken prion protein by in situ hybridization, Proc. Natl Acad. Sci. USA 90 (1993) 4309–4313. [36] A.D. Steele, S. Lindquist, A. Aguzzi, The prion protein knockout mouse: a phenotype under challenge, Prion 1 (2007) 83–93. [37] C. Kuwahara, A.M. Takeuchi, T. Nishimura, K. Haraguchi, A. Kubosaki, Y. Matsumoto, K. Saeki, T. Yokoyama, S. Itohara, T. Onodera, Prions prevent neuronal cell-line death, Nature 400 (1999) 225–226. [38] D.R. Brown, W.J. Schulz-Schaeffer, B. Schmidt, H.A. Kretzschmar, Prion proteindeficient cells show altered response to oxidative stress due to decreased SOD-1 activity, Exp. Neurol. 146 (1997) 104–112. [39] C.E. Le Pichon, M.T. Valley, M. Polymenidou, A.T. Chesler, B.T. Sagdullaev, A. Aguzzi, S. Firestein, Olfactory behavior and physiology are disrupted in prion protein knockout mice, Nat. Neurosci. 12 (2009) 60–69. [40] D.R. Brown, K. Qin, J.W. Herms, A. Madlung, J. Manson, R. Strome, P.E. Fraser, T. Kruck, A. von Bohlen, W. Schulz-Schaeffer, A. Giese, D. Westaway, H. Kretzschmar, The cellular prion protein binds copper in vivo, Nature 390 (1997) 684–687. [41] A.D. Steele, J.G. Emsley, P.H. Ozdinler, S. Lindquist, J.D. Macklis, Prion protein (PrPc) positively regulates neural precursor proliferation during developmental and adult mammalian neurogenesis, Proc. Natl Acad. Sci. USA 103 (2006) 3416–3421. [42] S. Chen, A. Mange, L. Dong, S. Lehmann, M. Schachner, Prion protein as transinteracting partner for neurons is involved in neurite outgrowth and neuronal survival, Mol. Cell. Neurosci. 22 (2003) 227–233. [43] N.R. Cashman, R. Loertscher, J. Nalbantoglu, I. Shaw, R.J. Kascsak, D.C. Bolton, P.E. Bendheim, Cellular isoform of the scrapie agent protein participates in lymphocyte activation, Cell 61 (1990) 185–192. [44] C.C. Zhang, A.D. Steele, S. Lindquist, H.F. Lodish, Prion protein is expressed on long-term repopulating hematopoietic stem cells and is important for their selfrenewal, Proc. Natl Acad. Sci. USA 103 (2006) 2184–2189. [45] J. Collinge, M.A. Whittington, K.C. Sidle, C.J. Smith, M.S. Palmer, A.R. Clarke, J.G. Jefferys, Prion protein is necessary for normal synaptic function, Nature 370 (1994) 295–297. [46] S. Mouillet-Richard, M. Ermonval, C. Chebassier, J.L. Laplanche, S. Lehmann, J.M. Launay, O. Kellermann, Signal transduction through prion protein, Science 289 (2000) 1925–1928. [47] G. Schmitt-Ulms, G. Legname, M.A. Baldwin, H.L. Ball, N. Bradon, P.J. Bosque, K.L. Crossin, G.M. Edelman, S.J. DeArmond, F.E. Cohen, S.B. Prusiner, Binding of neural cell adhesion molecules (N-CAMs) to the cellular prion protein, J. Mol. Biol. 314 (2001) 1209–1225. [48] H. Bueler, A. Aguzzi, A. Sailer, R.A. Greiner, P. Autenried, M. Aguet, C. Weissmann, Mice devoid of PrP are resistant to scrapie, Cell 73 (1993) 1339–1347. [49] S. Brandner, S. Isenmann, A. Raeber, M. Fischer, A. Sailer, Y. Kobayashi, S. Marino, C. Weissmann, A. Aguzzi, Normal host prion protein necessary for scrapieinduced neurotoxicity, Nature 379 (1996) 339–343. [50] G. Mallucci, A. Dickinson, J. Linehan, P.C. Klohn, S. Brandner, J. Collinge, Depleting neuronal PrP in prion infection prevents disease and reverses spongiosis, Science 302 (2003) 871–874. [51] B. Chesebro, M. Trifilo, R. Race, K. Meade-White, C. Teng, R. LaCasse, L. Raymond, C. Favara, G. Baron, S. Priola, B. Caughey, E. Masliah, M. Oldstone, Anchorless prion protein results in infectious amyloid disease without clinical scrapie, Science 308 (2005) 1435–1439. [52] L. Westergard, H.M. Christensen, D.A. Harris, The cellular prion protein (PrP(C)): its physiological function and role in disease, Biochim. Biophys. Acta 1772 (2007) 629–644. E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 [53] C.I. Lasmezas, J.P. Deslys, R. Demaimay, K.T. Adjou, F. Lamoury, D. Dormont, O. Robain, J. Ironside, J.J. Hauw, BSE transmission to macaques, Nature 381 (1996) 743–744. [54] J. Collinge, K.C. Sidle, J. Meads, J. Ironside, A.F. Hill, Molecular analysis of prion strain variation and the aetiology of ‘new variant’ CJD, Nature 383 (1996) 685–690. [55] G. Streisinger, C. Walker, N. Dower, D. Knauber, F. Singer, Production of clones of homozygous diploid zebra fish (Brachydanio rerio), Nature 291 (1981) 293–296. [56] T.P. Barros, W.K. Alderton, H.M. Reynolds, A.G. Roach, S. Berghmans, Zebrafish: an emerging technology for in vivo pharmacological assessment to identify potential safety liabilities in early drug discovery, Br. J. Pharmacol. 154 (2008) 1400–1413. [57] C. Chakraborty, C.H. Hsu, Z.H. Wen, C.S. Lin, G. Agoramoorthy, Zebrafish: a complete animal model for in vivo drug discovery and development, Curr. Drug Metab. 10 (2009) 116–124. [58] A. Aguzzi, Cell biology: beyond the prion principle, Nature 459 (2009) 924–925. [59] R.B. Wickner, H.K. Edskes, F. Shewmaker, T. Nakayashiki, Prions of fungi: inherited structures and biological roles, Nat. Rev. Microbiol. 5 (2007) 611–618. [60] M.F. Tuite, Yeast prions and their prion-forming domain, Cell 100 (2000) 289–292. [61] J.M. Gabriel, B. Oesch, H. Kretzschmar, M. Scott, S.B. Prusiner, Molecular cloning of a candidate chicken prion protein, Proc. Natl Acad. Sci. USA 89 (1992) 9097–9101. [62] T. Simonic, S. Duga, B. Strumbo, R. Asselta, F. Ceciliani, S. Ronchi, cDNA cloning of turtle prion protein, FEBS Lett. 469 (2000) 33–38. [63] B. Strumbo, S. Ronchi, L.C. Bolis, T. Simonic, Molecular cloning of the cDNA coding for Xenopus laevis prion protein, FEBS Lett. 508 (2001) 170–174. [64] E. Rivera-Milla, C.A. Stuermer, E. Malaga-Trillo, An evolutionary basis for scrapie disease: identification of a fish prion mRNA, Trends Genet. 19 (2003) 72–75. [65] F. Wopfner, G. Weidenhofer, R. Schneider, A. von Brunn, S. Gilch, T.F. Schwarz, T. Werner, H.M. Schatzl, Analysis of 27 mammalian and 9 avian PrPs reveals high conservation of flexible regions of the prion protein, J. Mol. Biol. 289 (1999) 1163–1178. [66] T. van Rheede, M.M. Smolenaars, O. Madsen, W.W. de Jong, Molecular evolution of the mammalian prion protein, Mol. Biol. Evol. 20 (2003) 111–121. [67] E. Rivera-Milla, B. Oidtmann, C.H. Panagiotidis, M. Baier, T. Sklaviadis, R. Hoffmann, Y. Zhou, G.P. Solis, C.A. Stuermer, E. Malaga-Trillo, Disparate evolution of prion protein domains and the distinct origin of Doppel- and prion-related loci revealed by fish-to-mammal comparisons, FASEB J. 20 (2006) 317–319. [68] E. Cotto, M. Andre, J. Forgue, H.J. Fleury, P.J. Babin, Molecular characterization, phylogenetic relationships, and developmental expression patterns of prion genes in zebrafish (Danio rerio), FEBS J. 102 (3) (2005) 651–655. [69] L. Calzolai, D.A. Lysek, D.R. Perez, P. Guntert, K. Wuthrich, Prion protein NMR structures of chickens, turtles, and frogs, Proc. Natl Acad. Sci. USA 102 (3) (2005) 651–655. [70] I. Vorberg, M.H. Groschup, E. Pfaff, S.A. Priola, Multiple amino acid residues within the rabbit prion protein inhibit formation of its abnormal isoform, J. Virol. 77 (2003) 2003–2009. [71] Y. Wen, J. Li, W. Yao, M. Xiong, J. Hong, Y. Peng, G. Xiao, D. Lin, Unique structural characteristics of the rabbit prion protein, J. Biol. Chem. 285 (41) (2010) 31682–31693. [72] V. Beringue, J.L. Vilotte, H. Laude, Prion agent diversity and species barrier, Vet. Res. 39 (2008) 47. [73] A.F. Hill, J. Collinge, Species-barrier-independent prion replication in apparently resistant species, APMIS 110 (2002) 44–53. [74] S.B. Prusiner, M. Scott, D. Foster, K.M. Pan, D. Groth, C. Mirenda, M. Torchia, S.L. Yang, D. Serban, G.A. Carlson, et al., Transgenetic studies implicate interactions between homologous PrP isoforms in scrapie prion replication, Cell 63 (1990) 673–686. [75] M.M. Simmons, J. Spiropoulos, S.A. Hawkins, S.J. Bellworthy, S.C. Tongue, Approaches to investigating transmission of spongiform encephalopathies in domestic animals using BSE as an example, Vet. Res. 39 (2008) 34. [76] J. Collinge, A.R. Clarke, A general model of prion strains and their pathogenicity, Science 318 (2007) 930–936. [77] B. Sweeting, M.Q. Khan, A. Chakrabartty, E.F. Pai, Structural factors underlying the species barrier and susceptibility to infection in prion disease, Biochem. Cell Biol. 88 (2010) 195–202. [78] Y. Wang, B. Hu, Y. Liu, H. Zhang, Y. Song, New insight into the species barrier of mammalian prions obtained from structure-based conservation analysis, Intervirology 51 (2008) 210–216. [79] A.M. Thackray, M.A. Klein, R. Bujdoso, Subclinical prion disease induced by oral inoculation, J. Virol. 77 (2003) 7991–7998. [80] R.H. Kimberlin, C.A. Walker, Pathogenesis of mouse scrapie: effect of route of inoculation on infectivity titres and dose–response curves, J. Comp. Pathol. 88 (1978) 39–47. [81] C. Hoffmann, U. Ziegler, A. Buschmann, A. Weber, L. Kupfer, A. Oelschlegel, B. Hammerschmidt, M.H. Groschup, Prions spread via the autonomic nervous system from the gut to the central nervous system in cattle incubating bovine spongiform encephalopathy, J. Gen. Virol. 88 (2007) 1048–1055. [82] A.F. Hill, J. Collinge, Subclinical prion infection in humans and animals, Br. Med. Bull. 66 (2003) 161–170. [83] A.M. Thackray, M.A. Klein, A. Aguzzi, R. Bujdoso, Chronic subclinical prion disease induced by low-dose inoculum, J. Virol. 76 (2002) 2510–2517. [84] S.J. Collins, V. Lewis, M.W. Brazier, A.F. Hill, V.A. Lawson, G.M. Klug, C.L. Masters, Extended period of asymptomatic prion disease after low dose inoculation: assessment of detection methods and implications for infection control, Neurobiol. Dis. 20 (2005) 336–346. 413 [85] L. Favre-Krey, M. Theodoridou, E. Boukouvala, C.H. Panagiotidis, A.I. Papadopoulos, T. Sklaviadis, G. Krey, Molecular characterization of a cDNA from the gilthead sea bream (Sparus aurata) encoding a fish prion protein, Comp. Biochem. Physiol. B Biochem. Mol. Biol. 147 (2007) 566–573. [86] B. Oidtmann, D. Simon, N. Holtkamp, R. Hoffmann, M. Baier, Identification of cDNAs from Japanese pufferfish (Fugu rubripes) and Atlantic salmon (Salmo salar) coding for homologues to tetrapod prion proteins, FEBS Lett. 538 (2003) 96–100. [87] J.H. Postlethwait, Y.L. Yan, M.A. Gates, S. Horne, A. Amores, A. Brownlie, A. Donovan, E.S. Egan, A. Force, Z. Gong, C. Goutel, A. Fritz, R. Kelsh, E. Knapik, E. Liao, B. Paw, D. Ransom, A. Singer, M. Thomson, T.S. Abduljabbar, P. Yelick, D. Beier, J.S. Joly, D. Larhammar, F. Rosa, M. Westerfield, L.I. Zon, S.L. Johnson, W.S. Talbot, Vertebrate genome evolution and the zebrafish gene map, Nat. Genet. 18 (1998) 345–349. [88] M. Premzl, L. Sangiorgio, B. Strumbo, J.A. Marshall Graves, T. Simonic, J.E. Gready, Shadoo, a new protein highly conserved from fish to mammals and with similarity to prion protein, Gene 314 (2003) 89–102. [89] T. Suzuki, T. Kurokawa, H. Hashimoto, M. Sugiyama, cDNA sequence and tissue expression of Fugu rubripes prion protein-like: a candidate for the teleost orthologue of tetrapod PrPs, Biochem. Biophys. Res. Commun. 294 (2002) 912–917. [90] R. Young, B. Passet, M. Vilotte, E.P. Cribiu, V. Beringue, F. Le Provost, H. Laude, J.L. Vilotte, The prion or the related Shadoo protein is required for early mouse embryogenesis, FEBS Lett. 583 (2009) 3296–3300. [91] J.C. Watts, B. Drisaldi, V. Ng, J. Yang, B. Strome, P. Horne, M.S. Sy, L. Yoong, R. Young, P. Mastrangelo, C. Bergeron, P.E. Fraser, G.A. Carlson, H.T. Mount, G. Schmitt-Ulms, D. Westaway, The CNS glycoprotein Shadoo has PrP(C)-like protective properties and displays reduced levels in prion infections, EMBO J. 26 (2007) 4038–4050. [92] E. Málaga-Trillo, G.P. Solis, Y. Schrock, C. Geiss, L. Luncz, V. Thomanetz, C.A. Stuermer, Regulation of embryonic cell adhesion by the prion protein, PLoS Biol. 7 (2009) e1000055. [93] M. Miesbauer, T. Bamme, C. Riemer, B. Oidtmann, K.F. Winklhofer, M. Baier, J. Tatzelt, Prion protein-related proteins from zebrafish are complex glycosylated and contain a glycosylphosphatidylinositol anchor, Biochem. Biophys. Res. Commun. 341 (2006) 218–224. [94] A. Force, M. Lynch, F.B. Pickett, A. Amores, Y.L. Yan, J. Postlethwait, Preservation of duplicate genes by complementary, degenerative mutations, Genetics 151 (1999) 1531–1545. [95] A. Nasevicius, S.C. Ekker, Effective targeted gene ‘knockdown’ in zebrafish, Nat. Genet. 26 (2000) 216–220. [96] E. Malaga-Trillo, E. Sempou, PrPs: Proteins with a purpose (lessons from the zebrafish), Prion 3 (2009) 129–133. [97] R. Chiesa, D.A. Harris, Fishing for prion protein function, PLoS Biol. 7 (2009) e75. [98] M.C. Sorgato, C. Peggion, A. Bertoli, Is indeed the prion protein an Harlequin servant of “many” masters? Prion 3 (4) (2009) 202–205. [99] J.W. Wilesmith, G.A. Wells, M.P. Cranwell, J.B. Ryan, Bovine spongiform encephalopathy: epidemiological studies, Vet. Rec. 123 (1988) 638–644. [100] D. Matthews, B.C. Cooke, The potential for transmissible spongiform encephalopathies in non-ruminant livestock and fish, Rev. Sci. Tech. 22 (2003) 283–296. [101] S.L. Benestad, J.N. Arsac, W. Goldmann, M. Noremark, Atypical/Nor98 scrapie: properties of the agent, genetics, and epidemiology, Vet. Res. 39 (2008) 19. [102] T. Baron, A.G. Biacabe, J.N. Arsac, S. Benestad, M.H. Groschup, Atypical transmissible spongiform encephalopathies (TSEs) in ruminants, Vaccine 25 (2007) 5625–5630. [103] C. Casalone, G. Zanusso, P. Acutis, S. Ferrari, L. Capucci, F. Tagliavini, S. Monaco, M. Caramelli, Identification of a second bovine amyloidotic spongiform encephalopathy: molecular similarities with sporadic Creutzfeldt–Jakob disease, Proc. Natl Acad. Sci. USA 101 (2004) 3065–3070. [104] G.A. Wells, A.C. Scott, C.T. Johnson, R.F. Gunning, R.D. Hancock, M. Jeffrey, M. Dawson, R. Bradley, A novel progressive spongiform encephalopathy in cattle, Vet. Rec. 121 (1987) 419–420. [105] L. Ingrosso, B. Novoa, A.Z. Valle, F. Cardone, R. Aranguren, M. Sbriccoli, S. Bevivino, M. Iriti, Q. Liu, V. Vetrugno, M. Lu, F. Faoro, S. Ciappellano, A. Figueras, M. Pocchiari, Scrapie infectivity is quickly cleared in tissues of orally-infected farmed fish, BMC Vet. Res. 2 (2006) 21. [106] E. Salta, C. Panagiotidis, K. Teliousis, S. Petrakis, E. Eleftheriadis, F. Arapoglou, N. Grigoriadis, A. Nicolaou, E. Kaldrymidou, G. Krey, T. Sklaviadis, Evaluation of the possible transmission of BSE and scrapie to gilthead sea bream (Sparus aurata), PLoS ONE 4 (2009) e6175. [107] H. Budka, Histopathology and immunohistochemistry of human transmissible spongiform encephalopathies (TSEs), Arch. Virol. Suppl. (2000) 135–142. [108] U. Unterberger, T. Voigtlander, H. Budka, Pathogenesis of prion diseases, Acta Neuropathol. 109 (2005) 32–48. [109] W. Butler, Comparative Vertebrate Neuroanatomy: Evolution and Adaptation, Wiley-IEEE, 2005. [110] C. Salas, C. Broglio, E. Duran, A. Gomez, F.M. Ocana, F. Jimenez-Moya, F. Rodriguez, Neuropsychology of learning and memory in teleost fish, Zebrafish 3 (2006) 157–171. [111] G. Almer, J.A. Hainfellner, T. Brucke, K. Jellinger, R. Kleinert, G. Bayer, O. Windl, H. A. Kretzschmar, A. Hill, K. Sidle, J. Collinge, H. Budka, Fatal familial insomnia: a new Austrian family, Brain 122 (Pt 1) (1999) 5–16. [112] J. Castri, R. Thiery, J. Jeffroy, P. de Kinkelin, J.C. Raymond, Sea bream Sparus aurata, an asymptomatic contagious fish host for nodavirus, Dis. Aquat. Organ. 47 (2001) 33–38. [113] R. Aranguren, C. Tafalla, B. Novoa, A. Figueras, Experimental transmission of encephalopathy and retinopathy induced by nodavirus to sea bream, Sparus aurata L., using different infection models, J. Fish Dis. 25 (2002) 317–324. 414 E. Málaga-Trillo et al. / Biochimica et Biophysica Acta 1812 (2011) 402–414 [114] C.G. Becker, T. Becker, Adult zebrafish as a model for successful central nervous system regeneration, Restor. Neurol. Neurosci. 26 (2008) 71–80. [115] G.K.H. Zupanc, Towards brain repair: insights from teleost fish, Semin. Cell Dev. Biol. 20 (2009) 683–690. [116] Y. Ano, A. Sakudo, H. Nakayama, T. Onodera, Uptake and dynamics of infectious prion protein in the intestine, Protein Pept. Lett. 16 (2009) 247–255. [117] A.D. Vale, The professional phagocytes of sea bass (Dicentrarchus labrax L.): cytochemical characterization of neutrophils and macrophages in the normal and inflamed peritoneal cavity, Fish & Shelfish Immunology 13 (2002) 183–198. [118] M. Chabrillon, Interactions of fish pathogenic Vibrio species with the mucous surfaces of gilthead seabream (Sparus aurata), Recent Research Developments in Microbiology, 2003, pp. 41–53. [119] A.Z. Dalla Valle, M. Iriti, F. Faoro, C. Berti, S. Ciappellano, In vivo prion protein intestinal uptake in fish, APMIS 116 (2008) 173–180. [120] M. Beekes, P.A. McBride, The spread of prions through the body in naturally acquired transmissible spongiform encephalopathies, FEBS J. 274 (2007) 588–605. [121] A. Aguzzi, M. Heikenwalder, G. Miele, Progress and problems in the biology, diagnostics, and therapeutics of prion diseases, J. Clin. Invest. 114 (2004) 153–160. [122] J.C. Espinosa, M. Morales, J. Castilla, M. Rogers, J.M. Torres, Progression of prion infectivity in asymptomatic cattle after oral bovine spongiform encephalopathy challenge, J. Gen. Virol. 88 (2007) 1379–1383. [123] T. Hastein, A. Hellstrom, G. Jonsson, N.J. Olesen, E.R. Parnanen, Surveillance of fish diseases in the Nordic countries, Acta Vet. Scand. Suppl. 94 (2001) 43–50. [124] Opinion of the Scientific Panel on Biological Hazards on a request from the European Parliament on the assessment of the health risks of feeding of ruminants with fishmeal in relation to the risk of TSE, EFSA J. 443 (2007) 1–26. [125] J.D. Best, W.K. Alderton, Zebrafish: an in vivo model for the study of neurological diseases, Neuropsychiatr. Dis. Treat. 4 (2008) 567–576. [126] N. Fernandez-Borges, J. de Castro, J. Castilla, In vitro studies of the transmission barrier, Prion 3 (2009) 220–223. [127] G.P. Saborio, B. Permanne, C. Soto, Sensitive detection of pathological prion protein by cyclic amplification of protein misfolding, Nature 411 (2001) 810–813. [128] R.A. Armstrong, P.L. Lantos, N.J. Cairns, What determines the molecular composition of abnormal protein aggregates in neurodegenerative disease? Neuropathology 28 (2008) 351–365. Biochimica et Biophysica Acta 1812 (2011) 415–422 Contents lists available at ScienceDirect Biochimica et Biophysica Acta j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / b b a d i s Review Microdomain-forming proteins and the role of the reggies/flotillins during axon regeneration in zebrafish☆ Claudia A.O. Stuermer ⁎ University of Konstanz, Department of Biology, 78465 Konstanz, Germany a r t i c l e i n f o Article history: Received 15 September 2010 Received in revised form 30 November 2010 Accepted 2 December 2010 Available online 13 December 2010 Keywords: Axon regeneration Microdomain Reggie/flotillin Recruitment Targeted delivery a b s t r a c t The two proteins reggie-1 and reggie-2 (flotillins) were identified in axon-regenerating neurons in the central nervous system and shown to be essential for neurite growth and regeneration in fish and mammals. Reggies/ flotillins are microdomain scaffolding proteins sharing biochemical properties with lipid raft molecules, form clusters at the cytoplasmic face of the plasma membrane and interact with signaling molecules in a cell type specific manner. In this review, reggie microdomains, lipid rafts, related scaffolding proteins and caveolin— which, however, are responsible for their own microdomains and functions—are introduced. Moreover, the function of the reggies in axon growth is demonstrated: neurons fail to extend axons after reggie knockdown. Furthermore, our current concept of the molecular mechanism underlying reggie function is presented: the association of glycosyl-phophatidyl inositol (GPJ)-anchored surface proteins with reggie microdomains elicits signals which activate src tyrosine and mitogen-activated protein kinases, as well as small guanosine 5′triphosphate-hydrolyzing enzymes. This leads to the mobilization of intracellular vesicles and to the recruitment of bulk membrane and specific cargo proteins, such as cadherin, to specific sites of the plasma membrane such as the growth cone of elongating axons. Thus, reggies regulate the targeted delivery of cargo— a process which is required for process extension and growth. This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. © 2010 Elsevier B.V. All rights reserved. 1. Introduction The zebrafish and goldfish visual pathway has served as a model system for the exploration of rules governing over the development of the retinotopic projection of retinal ganglion cell (RGC) axons onto the optic tectum [1,2]. The restoration of the retinotopic map by regenerating RGC axons after optic nerve lesion has also been intensively analyzed in this system. Because of its remarkable regeneration-supporting properties, the fish optic nerve is, moreover, suited to search for factors underlying successful regeneration of injured CNS fiber tracts per se. It is expected that insights into conditions allowing successful regrowth in fish might provide valuable clues for stimulation of axon regeneration in the mammalian CNS where regeneration does not spontaneously occur to any significant extent. Experiments along these lines have led to the discovery of two proteins, reggie-1 and reggie-2 [3], which turned out to be essential for growth and regeneration [4], hence their name reggie. Evidence for a role of the reggies in axon growth and regeneration, together with our current understanding of the underlying molecular ☆ This article is part of a Special Issue entitled Zebrafish Models of Neurological Diseases. ⁎ Tel.: +49 7531 882236. E-mail address: [email protected]. 0925-4439/$ – see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.bbadis.2010.12.004 mechanisms, will be reviewed in this article. But before reggies are considered in the context of axon growth, their properties as constituents of “lipid rafts,” or better microdomains will be discussed and compared to other microdomains and microdomain-forming proteins in zebrafish. Prior to this, however, the reasons why axons in the CNS of fish regenerate—as opposed to mammals which do not— will be considered in brief. 2. Axon regeneration in the CNS and the identification of the reggie proteins One reason for successful regeneration in the CNS is the growthpermissive nature of the glial cell environment in the fish optic nerve [5,6]. The mammalian CNS, by contrast, presents several growth inhibiting molecules on glial cells and in the extracellular matrix which block axon growth and prevent regeneration [7]. The second reason for the success of axon regeneration is also dependent on the special neuron-intrinsic properties of nerve cells in fish [8]. Retinal ganglion cells (RGCs), for instance, promptly upregulate growthassociated proteins after optic nerve lesion which enables the neurons to regrow their axons and to provide them with cytoskeletal proteins, cell adhesion molecules, the necessary receptors for guidance cues and, of course, bulk membrane material [4,8,9]. This ability to switch on expression of genes which were active during development but downregulated thereafter is by far less developed in mammals 416 C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 although in principle possible [10 and see below]. More than a dozen growth-associated proteins in RGCs have been identified at the molecular level, and these were recently complemented by a proteomics approach in mammalian cortical neurons [11]. The two proteins discovered in this context, reggie-1 and reggie-2 [3,11], are 45% identical to one another and have a molecular weight of 48 kD. Reggies were also identified in mammals [12,13] where they occur in basically all cells examined so far. In the rat retina, they are upregulated in exactly the 3% of RGCs that, as demonstrated by Richardson et al. [10], can regenerate their axon in a growthpermissive environment. However, it was not immediately clear how reggies regulate regeneration. Unlike cell adhesion molecules which are also upregulated during axon regeneration [14,15] and mediate the interaction of the elongating growth cone with the environment, the reggies were found to reside at the cytoplasmic face of the plasma membrane. They are anchored to the membrane by myristoyl and palmitoyl residues [16,17] and through stretches of hydrophobic amino acids in their N-terminal head domain (Fig. 1). It was also observed that the reggies do not only occur at the plasma membrane but also in association with many intracellular vesicles [9,18,19], including the Rab11 recycling compartment. Consistent with their identity as microdomain proteins, reggies were independently identified as components of the protein fraction known as DRMs (detergent resistant membranes). The DRM contains molecules that are insoluble in non-ionic detergents (such as Triton X-100) and that are regarded as “lipid raft” components. Together with other DRM proteins, the reggies “float” after sucrose density centrifugation [12], and hence, they were named flotillins [12]. Reggie-1 corresponds to flotillin-2, and reggie-2 to flotillin-1. Reggies/flotillins are often used as markers for the lipid raft fraction in biochemical work. We prefer the term “reggie microdomain” (instead of lipid raft) because reggies assemble by oligomerization into clusters of ≤100 nm [9,20] and are, therefore, visible at the light microscopic level by immunofluorescence analysis [13,20,21] with specific antibodies. The microdomains appear as dots or puncta at the plasma membrane of cells and along axons, growth cones, filopodia and as densely packed rows of dots, for instance, at cell–cell contacts [21,22] and in the T cell cap [20,23]. This “pearl on a string-like” arrangement in T cell caps and at cell contacts (Fig. 2) is formed by a reduction of the distance between individual microdomains [22]. Fig. 1. The microdomain scaffolding proteins and their domains. Reggie-1 and reggie-2 (flotillin) possess an N-terminal head domain (light blue), with two hydrophobic regions (yellowish) and acylation sites (red) for membrane attachment. The tail domain (gray) has strikingly many EA repeats for α-helical coiled-coil formation and oligomerization. The other microdomain scaffolding proteins share some degree of sequence homology with reggie and among each other and also form oligomers. C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 417 Fig. 2. Distribution and function of reggie microdomains in T cells and neurons. A) In T cells, reggie-1 (and -2)—visualized here by a reggie-1 specific antibody (red)—is clustered at one pole of the cell, the preformed reggie cap which represents a platform for the association of specific GPI-anchored proteins, such as PrP and Thy-1 and signaling. B) Reggie microdomains cluster at cell–cell contact sites (arrowheads) by reducing the distance between individual microdomains (red, immunofluorescence with anti-reggie-1 antibodies). C) The growth cone from a zebrafish RGC axon shows the distribution of reggie microdomains (puncta) at its central region as well as in its filopodia. D) Individual zebrafish RGC neuron and its axon immunostained with a reggie-1 specific antibody and showing the punctate microdomains along the axon. E) Hippocampal neurons differentiate in vitro (E, control siRNA) and extend neurites whereas their counterparts fail to produce neurites after exposure to reggie-1 specific siRNAs (F). G) Hippocampal neurons often produce club-shaped endings where a growth cone should have been, after perturbation of reggie-1 expression. Scale bars: 20 μm. 3. Microdomains and lipid rafts Microdomains, caveolae and lipid rafts represent functional units in cellular membranes whose existence depends on mainly saturated sphingolipids and cholesterol in a liquid ordered phase and segregated from unsaturated lipids in the liquid disordered phase [24,25]. Rafts range in size from a few nanometer encompassing by estimate six GPI-anchored proteins (nanoclusters) [26] to several hundred nanometers (known as “clustered rafts”) [25] are dynamic and generally short-lived. They can be stabilized by the actin cytoskeleton [27] which influences their distribution, half life and size. The existence of lipid rafts has been heavily disputed [28] particularly since they often escape microscopic observation due to their small size [26]. Recent results now demonstrate the phase separation of sphingolipids and cholesterol in giant plasma membrane vesicles of vertebrate cells which support again the raft concept [29]. Microdomains—in contrast to rafts—depend on specific scaffolding proteins that have an affinity for the lipid raft environment and share many raft properties. A family of microdomain proteins with a characteristic membrane interaction domain at their N-terminal “head” domain consists of stomatin, prohibitin, flotillin/reggie, podocin and erlin [9,30–32]. The scaffolding proteins (Fig. 1) possess a region for membrane interaction and stretches of amino acids (tail domain) which promote oligomerization [33]. These protein oligomers, when labeled by antibodies or tags, make microdomains visible at the light microscopic level and amenable to direct analysis of distribution and function. Lipid rafts and microdomains confer inhomogeneities to the plasma membrane and allow the formation of membrane compartments [25] which are used by specific membrane proteins for cluster formation, protein–protein interactions and signaling [34] as well as for the recruitment and targeted delivery of membrane proteins to specific sites [9]. GPI-anchored proteins are typical cell-surface residents of rafts and microdomains whereas proteins with acyl residues (palmitoylation, myristoylation and others) typically reside at the cytoplasmic face of the microdomain plasma membrane. Certain transmembrane proteins also acquire a preference for the lipid raft and microdomain environment, in particular certain growth factor and transmitter receptors [25]. While being intensively discussed among cell biologist and experts in lipidomics, rafts and microdomains do not play such a central role in research on zebrafish development or disease. Few exceptions are the microdomains scaffolded by the proteins caveolin (i.e. caveolae), podocin and reggie which will be described next and whose function in zebrafish embryos and/or adult fish will be discussed subsequently. 418 C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 4. Microdomains and caveolae Among the most intensively studied microdomains are caveolae, 100 nm flask-shaped plasma membrane invaginations, whose formation depends on caveolin [35]. Each of the other scaffolding proteins, reggie/flotillin, stomatin, podocin, prohibitin and erlin, forms its own microdomain without producing membrane indentations or other morphologically conspicuous structures. There was a debate as to whether reggies represent additional constituents of caveolae which would imply that they serve as a second scaffold of the caveolar invaginations [12]. Reggie proteins, however, are present in basically every cell analyzed so far, whereas caveolin is not. Caveolin, not reggie, is absolutely required for caveola formation. Moreover, caveolae are absent from neurons and lymphocytes [13,20,21,36] which do have reggie microdomains. Even in cells with caveolae, such as astrocytes, caveolae and reggie proteins occupy non-overlapping regions of the plasma membrane [21], so that reggies are clearly not constituents of caveolae. In fact, we are not aware of any report on microdomains with mixed protein scaffolds. The microdomain scaffolding proteins are evolutionarily highly conserved [32,37] suggesting important functions across kingdoms of life. Accordingly, inspection of databases shows the presence of homologs of caveolin, stomatin, prohibitin, reggie/flotillin, podocin and erlin in zebrafish. Aside from their sequence, however, not much is known on their function. We found no report on zebrafish stomatin, prohibitin and erlin. But work was published on caveolin-1 and caveolin-3 in the early embryo, on podocin during kidney formation, and on reggie-1 and reggie-2 during axon regeneration. The reggies will be considered in greater detail because they were discovered in the fish CNS and functionally characterized in the context of axon regeneration [3,4]. Moreover, results from fish together with insights from various cell types and species has led to a new concept explaining reggie functions [9] but before its presentation published data on zebrafish caveolin-1, caveolin-3 and podocin will be summarized. 5. Zebrafish caveolin and caveolae In mammals, caveolin and caveolae have been implicated in many functions such as endocytosis, transcytosis, signaling, trafficking of cholesterol (for review see reference 38) so that it came as a surprise that the caveolin-1 knockout mice had at first sight no apparent phenotype. But mice turned out to suffer from pulmonary defects, particularly under the challenge of exercise, and to lack the ability to appropriately adapt their blood pressure [39]. 5.1. Caveolin-1 In zebrafish, caveolin-1 and caveolae were shown to exist during the development in neuromasts and in notochord cells (and other structures) [40] which exhibited caveolae in abundance. Morpholino-mediated knockdown to assess the potential role of caveolin in embryogenesis showed that the formation of neuromasts was perturbed and the notochord disrupted [40]. Another publication reports additional defects in the eye and in somites at 12 h post fertilization (hpf) as well as disruption of the actin cytoskeleton [41]. Later defects emerged in the vascular endothelium [41] and in the smaller size of the embryos [40]. In addition, the lateral line developed abnormally most likely because of the defects in neuromasts and notochord [40,41]. The embryos might suffer from other problems as, for instance, reduced cholesterol levels of the plasma membrane since caveolin has been proposed to function as a carrier for cholesterol [42] and cholesterol is needed for the existence of lipid rafts and microdomains [25]. Whether morpholino-treated fish develop severe abnormalities due to imbalanced cholesterol levels has not been reported nor is clear how caveolindependent receptor-ligand interaction and signal transduction may change. The enzymes involved in cholesterol synthesis [43] are expressed in zebrafish and in the embryo as maternal messages at the one and two cell stage predicting that cholesterol biosynthesis is crucial already in the early blastocytes. Altogether, it appears that caveolin-1 and caveolae control the differentiation of specific cell types in the embryo, yet the underlying mechanism is not well understood. 5.2. Caveolin-3 Zebrafish also expresses a homolog of mammalian caveolin-3 (cav-3) which is a muscle-specific isoform in mammals and fish [44,45]. Cav-3 in zebrafish was found in the first differentiating muscle precursor cells and later in skeletal and cardiac muscle fibers where expression correlated with the presence of caveolae. Cav-3, however, also appeared in pectoral fin, facial muscle and notochord [44]. Downregulation of cav-3 expression caused severe muscle abnormalities and uncoordinated movement and increased the number of muscle pioneer-like cells adjacent to the notochord. The signaling pathways which depend on cav-3 and caveolae and steer muscle development have not been described and are to this end unknown. 6. Zebrafish podocin Podocin is expressed by podocytes that are necessary for the blood filtration barrier in the kidney glomerulus [46,47]. Podocytes form the slit diaphragm, a specialized cell–cell adhesion complex at the podocyte foot processes surrounding glomerular blood vessels. Slit diaphragm formation depends on the cell adhesion proteins (of the immunoglobulin superfamily) nephrin and Neph as well as podocin [48]. Much as in higher vertebrates, zebrafish nephrin and podocin are specifically expressed in pronephric podocytes and required for the development of pronephric podocyte cell structures [49]. Morpholino-mediated knockdown of podocin (or nephrin) resulted in loss of a slit diaphragm at 72 and 96 hpf and failure to form normal podocyte foot processes. Absence of normal podocin (or nephrin) expression resulted in defects in glomerular filtration and aberrant passage of high molecular filtrate [49,50]. The exact cellular and molecular mechanism of podocin function during slit diaphragm formation is not known. However, human podocin mutations Cterminal to the membrane association (head) domain resulted in loss of podocin at the cell membrane and aberrant accumulation in ER secretory pathway vesicles [51]. Podocin mutants also resulted in failed delivery of nephrin to the slit diaphragm thus indicating a role of podocin in nephrin trafficking [48]. 7. Reggie microdomains Reggie microdomains are defined by hetero-oligomeric clusters of reggie-1 and reggie-2. To this end, it has not been explicitly examined whether reggie-1 and reggie-2 can perform separate functions. However, it has become clear that reggie-2 undergoes proteasomal degradation when reggie-1 is downregulated (53), for example after siRNA and morpholino-mediated interference with reggie-1 expression levels in neurons, epithelial cells and zebrafish embryo. It is unclear why reggie-2 is so tightly regulated in dependence of reggie-1 nor has it been thoroughly analyzed whether reggie-1 and reggie-2 can subserve separate functions in other cell types. Therefore, we do not distinguish between reggie-1 and -2 in the text when they were not addressed individually. C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 7.1. Association of the GPI-anchored prion protein and Thy-1 with reggie microdomains Reggie-1 and -2 were isolated from larval goldfish by coimmunoprecipitation with an antibody against goldfish Thy-1 [3]. This is in line with the fact that Thy-1, as a GPI-anchored protein, associates preferentially with reggie microdomains [20,21,52,53] not only in fish but also in all vertebrate cells studied so far. Moreover, Thy-1 is upregulated in fish RGCs after nerve lesion like reggie [53] and co-localizes with reggie in microdomains in growth cones, along axons, at cell–cell contact sites and in the T cell cap [9]. In the uninjured fish visual system, Thy-1 as well as reggie, is colocalized in the few axons from new RGCs that are constantly added to the fish retina [3]. Thus, Thy-1 seems to cocluster preferentially with reggie microdomains in vitro and in vivo. This applies also to the GPIanchored cellular prion protein [PrPc; 20] and the zebrafish homologs PrP-1 and PrP-2 [54]. That this association of PrP and Thy-1 with reggie-1 and -2 is functionally important was demonstrated first in T cells as will be described below. Further insights into the role of the reggies came from studies on insulin stimulation of adipocytes (fat cells) [55,56] and on cell–cell contact formation in various types of cells [9]. 7.2. Reggie and PrP signaling in T cells In T cells, reggie microdomains are pre-clustered at one pole of the cell, representing ‘the preformed reggie cap’ [23] where the T cell receptor complex and the signaling molecules associated with T cell activation coalesce upon stimulation [20]. The preformed reggie cap is important for T cell activation. A dominant-negative reggie construct disrupts the cap and interferes with the formation of the immunological synapse [19]. In T cells, the activation of GPI-anchored proteins by antibody cross-linking is sufficient for TCR complex assembly in the cap [34]. This applies, for instance, to Thy-1 and PrP, whose activation leads to their selective association with the reggie cap [20,21]. When clustered in the reggie cap, PrP elicits the phosphorylation of the MAP (mitogen-activated protein) kinase ERK1/2 and evokes a distinct Ca2+ signal leading to the recruitment of CD3 [20,57], the major T cell receptor component, to the membrane in the cap. Yet, CD3 recruitment is not sufficient for the full activation of the T cell receptor complex [20], which would require cross-linking of CD3 or stimulation by antigen. T cell receptor recruitment involves the communication with actin dynamics through src tyrosine kinases (fyn, lck, src), Rho-GTPases and the signaling molecule Vav [19]. The conclusion from these results is that the microdomains consisting of reggie oligomers serve as platforms for PrP cluster formation, signaling, actin rearrangement and the recruitment of CD3 (from internal stores) to the cap [9]. The involvement of reggie in the recruitment of a membrane protein (Glut4) to specific sites of the plasma membrane has previously been observed in adipocytes [55,56] as is discussed next. 419 It is known that the small GTPases regulate actin dynamics and vesicle trafficking [4,60–62]. Vesicles carrying reggie were observed to shuttle between the plasma membrane and more distant intracellular sites [61] including the recycling compartment. This indicates that reggie microdomains at vesicle membranes participate in the signaling required for vesicle mobilization, trafficking and targeting. These events require the activity of small GTPases that, in turn, have a preference for rafts [9,58,63]. The reggie/flotillin-binding protein CAP/ponsin and its relatives ArgBP2 and vinexin were implied in the interaction with the actin cytoskeleton [56] in which reggie/flotillin is also involved [64] and with substrate and cell adhesion proteins, such as integrins and cadherins. Together, these findings suggest that reggie activates a pathway for polarized delivery of cargo [9]. The type of cargo depends on the cell type and includes, for example, transporters (such as Glut4) [55], cell adhesion molecules (such as cadherins, integrins) [54,56], receptors, ion channels and many other proteins that need to be targeted to specific sites of cells [9]. The sites are “marked” by the clustering of reggie/flotillin microdomains and co-clusters of GPIanchored proteins, which provides external signals and specificity (Fig. 3). Reggies/flotillins at membranes of intracellular vesicles [18,61] might represent similar signaling centers for clustering of signaling molecules to promote vesicle transport, sorting and targeted delivery of cargo such as cadherin (Fig. 3) [9]. 7.4. Reggies, PrP and cadherin recruitment Reggies are clustered at contact sites between cells [20,21] and so is PrP. PrP–PrP trans-interactions in contacts between mammalian 7.3. Reggie-associated signaling and the recruitment of Glut4 in adipocytes In adipocytes (fat cells) flotillin-1/reggie-2 interact with the c-cblassociated adaptor protein (CAP) and via CAP with a signaling complex that activates the small guanosine 5′-triphosphate-hydrolyzing enzyme (GTPase) TC10 [55] in lipid rafts (reggie/flotillin microdomains). This results in the translocation of the glucose transporter Glut4 from a specific vesicular storage compartment to the plasma membrane [55,56]. This step requires the exocyst and the (Ras-related) GTPase RalA. TC10 and RalA associate with vesicles, such as the recycling compartment and (transport) vesicles en route to the plasma membrane [58,59] which also carry reggie—in agreement with the role of the reggie proteins in activation of TC10 and other small GTPases. Fig. 3. Recruitment and targeted delivery of membrane and membrane proteins regulated by reggie. PrP is typically associated with reggie microdomains and clustered at contact sites between cells (cell 1 with cell 2). PrP trans-interaction in reggie microdomains leads to signal transduction including the CAP/reggie-associated signaling complex which activates the small GTPase TC10. This leads to the mobilization of cargo (cadherin)-loaded vesicles which also have reggie microdomains, through TC10 and RalA. Vesicles are transported (along cytoskeletal elements) to the plasma membrane and fuse resulting in membrane growth and cadherin delivery at cell contact sites. 420 C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 epithelial and neuronal cells or zebrafish embryonic blastocytes cause PrP clustering and co-clustering with reggies [9,54]. PrP clusters in reggie microdomains apparently regulate cell contact formation in zebrafish embryos as well as in mammalian cells (Fig. 3) as well as downstream actin rearrangement [54]. In zebrafish embryos, PrP-1 accumulates selectively at contact sites between cells at the blastula and gastrula stage. PrP clustering results in signaling. This signal is necessary for the PrP-dependent recruitment of epithelial (E)cadherin from the internal vesicle pool to the contacts between cells in zebrafish as well as in various mammalian cells [9,54]. In zebrafish, PrP downregulation by morpholinos has dramatic effects: the cells lose contact and the embryo dies from its failure to proceed through gastrulation [54]. Search for the underlying mechanism showed that E-cadherin was retained in intracellular recycling vesicles instead of emerging at the cell-surface where E-cadherin should be engaged in homophilic adhesion. How GPI-anchored proteins such as PrP communicate with reggie for the recruitment of cadherins and other membrane proteins remains to be solved. How proteins with lipid anchors confined to one leaflet of the lipid bilayer are capable of signal transduction without associated transmembrane proteins is also unsolved and debated in many reports and reviews dealing with lipid rafts and their function [25]. Recent results from a biophysical simulation approach, however, suggest that cluster formation of GPI-anchored proteins with lipid anchors in the outer leaflet of the plasma membrane is crucial and efficient in influencing the adjacent cytoplasmic leaflet, which offers another cluster of proteins (such as the reggie microdomain) with their own lipid anchors—the myristoyl and palmitoyl residues in reggie (Matthias Weiss, pers. commun.). Moreover, sphingolipids with lipid chains longer than the width of one leaflet are enriched in rafts and microdomains. It is thus conceivable that co-clusters of proteins in the outer and inner leaflet of the plasma membrane can transduce signals into the cell, activating src tyrosine kinases that are also lipid-anchored and able to interact with reggie (such as fyn) [65]. Thus, reggies appear to function as platforms necessary for the assembly of specific cell-surface proteins, for signal transduction and activation of small GTPases to regulate actin dynamics. Specificity and the trigger for signaling appear to come from the cell-surface associates of reggie/flotillin microdomains, which are cell type dependent (insulin receptor in adipocytes, PrP (Thy-1) in T cells and PrP (Thy-1) at cell contact sites in the embryo and cells in vitro) and domain specific (the cap in T cells, contact sites). The cell-surface-derived signals often (or always) target GTPases and actin dynamics, conceivably to activate and recruit specific vesicles from the intracellular pool such as Glut4 and E-cadherin. Thus, consistent with finding reggie-1 and -2 in basically all cell types and species as distant as flies, fish and mammals [53], their function seems to be of general importance for the communication between cells and for membrane protein sorting, trafficking and delivery. How does this relate to axon growth? 8. Reggies regulate axon regeneration To examine whether and how reggies might regulate axon regeneration and growth, we applied morpholinos in a piece of gelfoam to the transected right nerve of adult zebrafish and a control morpholino to the left transected nerve [4] following a procedure by Becker et al. [66]. The morpholinos are being taken up by the severed axons and transported retrogradely to the RGCs of origin. The morpholinos are labeled by a fluorescent tag (lissamine) thus allowing identification of morpholino-laden RGCs in retina whole mounts. Miniexplants of retinae of treated fish send out axons. The quantification of axons extending from the explants showed a 45% decrease from morpholino-treated RGCs compared to control. A second experiment gave even more striking effects: a fluorescent dye was applied to the regenerating axons behind the first transection and morpholino application site in the optic nerve which allows to quantify the regenerating axons by the second label in their parent RGCs. This approach showed a 70% reduction of RGCs with regenerating axons in the optic nerve showing that reggie morpholinos block axon regeneration and, in other words, suggest that reggies are necessary for axon growth and regeneration. This conclusion was supported by individual neurons from the mouse hippocampus in culture in which reggie expression levels were downregulated by reggie-specific siRNAs [4]. These neurons failed to differentiate and did not form axons or failed to elongate them. Axons often showed immobile club-shaped endings where a growth cone should have been and did not grow (Fig. 2). Other neurons had shorter or no processes (Fig. 2), were abnormally large and malformed and produced conspicuous bulges instead of axons and dendrites. The elongation process was apparently blocked because, as we believe, membrane proteins or membrane and proteins from the internal vesicle pool were not appropriately supplied to the prospectively growing tips [9]. Thus, the mechanism of membrane and protein recruitment which is necessary for process growth is controlled by reggies. Therefore, reggies clearly regulate regeneration and axon growth. Such phenotypes indicate a disturbance in the communication with the cytoskeleton, which was examined in N2a neuroblastoma cells. Reggie mis- and downregulation perturbed the activation levels of the Rho-GTPases Rac, Rho and cdc42 which changed the activation of the downstream effectors WASP, Arp2/3, cortactin and cofilin and of focal adhesion kinase (FAK), p38 and ras [4,62]. This is compatible with the view that the reggie proteins can affect and modify actin dynamics [64], in connection with the recruitment of membrane and proteins (or simply membrane building blocks) from the constitutive secretory pathway and recycling compartment to the growing tips [9]. 9. An explanation of reggie functions Taken together, it seems that reggies are crucial for the recruitment of membrane and specific membrane proteins (or membrane building blocks) to specific regions or domains of the cell (Fig. 3). Neurons with elongating and regenerating axons, where reggies are enriched at growing tips, seem to suffer badly when reggies are missing since blockage of reggie function impairs the delivery of material for growth. Membrane proteins are recruited from vesicular pools (recycling compartment, transport vesicles, Golgi-associated vesicles) and moved along cytoskeletal elements, and this requires the small GTPases of the Rho- and Ras-family and their effectors. Reggie-associated signal transduction and membrane protein recruitment appear to be activated by surface molecules with specific affinities for the reggie microdomain environment such as GPI-anchored proteins and certain receptor types. The surface proteins can be activated by ligands, binding partners (PrP–PrP trans-interaction) or cross-linking antibodies mimicking ligands which activate signaling molecules located in reggie microdomains: src tyrosine kinases, small GTPases, cbl and interacting adaptor complexes. Signaling provokes the recruitment of vesicles and thereby the targeted delivery of membrane building blocks and proteins for axon growth, adhesion, navigation as well as for contact formation between cells and for the establishment of specific membrane domains such as the leading edge, growth cone, its lamellipodia and filopodia. This implies a role of the reggies in the recruitment of specific vesicles and certain membrane proteins from vesicles and stores (transporters, receptors, ion channels and adhesion molecules) [9]. This role of the reggies in the mediation of the recruitment of proteins/membrane from intracellular pools requires their existence in all cells of vertebrates and invertebrates and explains why they are so important for axon growth. Upregulation of reggie expression is a C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 prerequisite for axon growth/regeneration not only in fish but also in the mammalian CNS [4,13]. In fish, all RGCs express reggie at high levels and regenerate their axons. In the mammalian retina, only 3% of the neurons regenerate their axons and express reggie. When new methods are applied to stimulate the remaining neurons to upregulate reggie, axon regeneration might be successful to an extent that would allow recovery of function in mammals. First results in this direction are encouraging (unpublished results: Jan Koch, Paul Lingor, Matthias Bähr [Göttingen] in collaboration with Gonzalo Solis and C.A.O. Stuermer [Konstanz]). In light of such results it is tempting to speculate that upregulation of reggie in injured neurons after spinal cord lesion might promote axon regrowth and improve the fatal conditions after lesion. Upregulation of reggie and the ensuing growth response might provide neurons with the necessary “power” to resist and counteract effects from neurodegenerative diseases. This unexpected role of the reggies predicts on the one hand related functions of other microdomain proteins and requires on the other hand identification of all players in the reggie-dependent signaling pathway. Moreover, whether reggies act as molecular shuttles between plasma membrane and intracellular vesicle pools needs to be addressed experimentally. The most crucial issue in the context of axonal regeneration is the identification of the factors that control reggie upregulation in neurons during axon outgrowth and regeneration in fish and in mammals so to stimulate neurons to produce new processes upon injury and resist degeneration. Acknowledgment Our work is supported by the Deutsche Forschungsgemeinschaft and Fonds der Chemischen Industrie. I want to thank Ulrike Binkle, Marianne Wiechers and Vsevolod Bodrikov for the figures and all co-workers for the experiments and results in our reggie research. References [1] R.M. Gaze, The Formation of Nerve Connections, Academic Press, London, 1970. [2] C.A.O. Stuermer, Retinotopic organization of the developing retinotectal projection in the zebrafish embryo, J. Neurosci. 8 (1988) 4513–4530. [3] T. Schulte, K.A. Paschke, U. Laessing, F. Lottspeich, C.A.O. Stuermer, Reggie-1 and reggie-2, two cell surface proteins expressed by retinal ganglion cells during axon regeneration, Development 124 (1997) 577–587. [4] C. Munderloh, G.P. Solis, F. Jaeger, M. Wiechers, C.A.O. Stuermer, Reggies/flotillins regulate retinal axon regeneration in the zebrafish optic nerve and signal transduction in N2a cells, J. Neurosci. 29 (2009) 6607–6615. [5] H.B Abdesselem, A. Shypitsyna, C.A.O. Stuermer, No Nogo66-mediated inhibition of regenerating axons in the zebrafish optic nerve, J. Neurosci. 29 (2009) 15489–15498. [6] H. Diekmann, M. Klinger, T. Oertle, M. Pogoda, D. Heinz, M.E. Schwab, C.A.O. Stuermer, Evidence for the absence of the neurite growth inhibitory protein Nogo-A in fish, J. Mol. Evol. 22 (2005) 1635–1648. [7] I.C. Maier, M.E. Schwab, Sprouting, regeneration and circuit formation in the injured spinal cord: factors and activity, Phil. Trans. R. Soc. B Biol. Sci. 361 (2006) 1611–1634. [8] C.A.O. Stuermer, M. Bastmeyer, M. Bähr, G. Strobel, K. Paschke, Trying to understand axonal regeneration in the CNS of fish, J. Neurobiol. 23 (1992) 537–550. [9] C.A.O. Stuermer, The reggie/flotillin connection to growth, Trends Cell Biol. (2009), doi:10.1016/j.tcb2009.10.003. [10] P.M. Richardson, U.M. McGuinness, A.J. Aguayo, Axons from CNS neurons regenerate into PNS grafts, Nature 284 (1980) 264–265. [11] A.M. Taylor, N.C. Berchtold, V.M. Perreau, C.H. Tu, N.L. Jeon, C.W. Cotman, Axonal mRNA in uninjured and regenerating cortical mammalian axons, J. Neurosci. 29 (2009) 4697–4707. [12] P.E. Bickel, P.E. Scherer, J.E. Schnitzer, P. Oh, M.P. Lisanti, H.F. Lodish, Flotillin and epidermal surface antigen define a new family of caveolae-associated integral membrane proteins, J. Biol. Chem. 272 (1997) 13793–13802. [13] D.M. Lang, S. Lommel, M. Jung, R. Ankerhold, B. Petrausch, U. Laessing, M.F. Wiechers, H. Plattner, C.A.O. Stuermer, Identification of reggie-1 and reggie-2 as plasmamembrane-associated proteins which cocluster with activated GPIanchored cell adhesion molecules in non-caveolar micropatches in neurons, J. Neurobiol. 37 (1998) 502–523. [14] K.A. Paschke, F. Lottspeich, C.A.O. Stuermer, Neurolin, a cell surface glycoprotein on growing retinal axons in the goldfish visual system, is reexpressed during retinal axonal regeneration, J. Cell Biol. 117 (1992) 863–875. 421 [15] C.A.O. Stuermer, C.A. Leppert, Molecular determinants of retinal axon pathfinding in fish, in: N.A. Ingoglia, M. Murray (Eds.), Axonal Regeneration in the Central Nervous System, Marcel Dekker Inc., New York, 2000. [16] I.C. Morrow, S. Rea, S. Martin, I.A. Prior, R. Prohaska, J.F. Hancock, D.E. James, R.G. Parton, Flotillin-1/reggie-2 traffics to surface raft domains via a novel Golgiindependent pathway. Identification of a novel membrane targeting domain and a role for palmitoylation, J. Biol. Chem. 277 (2002) 48834–48841. [17] C. Neumann-Giesen, I. Fernow, M. Amaddii, R. Tikkanen, Role of EGF-induced tyrosine phosphorylation of reggie-1/flotillin-2 in cell spreading and signaling to the actin cytoskeleton, J. Cell Sci. 120 (2007) 395–406. [18] J.F. Dermine, S. Duclos, J. Garin, F. St-Louis, S. Rea, R.G. Parton, M. Desjardins, Flotillin-1-enriched lipid raft domains accumulate on maturing phagosomes, J. Biol. Chem. 276 (2001) 18507–18512. [19] M. Langhorst, A. Reuter, G. Luxenhofer, E.M. Boneberg, D.F. Legler, H. Plattner, C.A.O. Stuermer, Preformed reggie/flotillin caps: stable priming platforms for macrodomain assembly in T cells, FASEB J. 20 (2006) 711–713. [20] C.A.O. Stuermer, M. Langhorst, M. Wiechers, D.F. Legler, S. Hannbeck von Hanwehr, A.H. Guse, H. Plattner, PrPc capping in T cells promotes its association with the lipid raft proteins reggie-1 and reggie-2 and leads to signal transduction, FASEB J. 2 (2004) 1–27. [21] C.A.O. Stuermer, D.M. Lang, F. Kirsch, M.F. Wiechers, S.-O. Deininger, H. Plattner, Glycosylphosphatidyl inositol-anchored proteins and fyn kinase assemble in noncaveolar plasma membrane microdomains defined by reggie-1 and -2, Mol. Biol. Cell 12 (2001) 3031–3045. [22] C.A.O. Stuermer, H. Plattner, The “lipid raft”/microdomain proteins reggie-1 and reggie-2 (flotillins) are scaffolds for protein interaction and signaling, Biochem. Soc. Symp. 72 (2005) 109–118. [23] L. Rajendran, M. Masilamani, S. Solomon, R. Tikkanen, C.A.O. Stuermer, Asymmetric localization of flotillins/reggies in preassembled platforms confers inherent polarity to hematopoietic cells, Proc. Natl Acad. Sci. USA 100 (2003) 8241–8246. [24] K. Simons, E. Ikonen, Functional rafts in cell membranes, Nature 387 (1997) 569–572. [25] K. Simons, W.L. Vaz, Model systems, lipid rafts, and cell membranes, Annu. Rev. Biophys. Biomol. Struct. 33 (2004) 269–295. [26] D. Goswami, K. Gowrishankar, S. Bilgrami, S. Ghosh, R. Raghupathy, R. Chadda, R. Vishwakarma, M. Rao, S. Mayor, Nanoclusters of GPI-anchored proteins are formed by cortical actin-driven activity, Cell 135 (20098) 1085–1097. [27] P.F. Lenne, L. Wawrezinieck, F. Conchonaud, O. Wurtz, A. Boned, X.-J. Guo, H. Rigneault, H.T. He, D. Marguet, Dynamic molecular confinement in the plasma membrane by microdomains and the cytoskeleton network, EMBO J. 25 (2006) 3245–3256. [28] J.F. Hancock, Lipid rafts: contentious only from simplistic standpoints, Nat. Rev. Mol. Cell Biol. 7 (2006) 456–462. [29] H.-J. Kaiser, D. Lingwood, I. Levental, J.L. Sampaio, L. Kalvodova, L. Rajendran, K. Simons, Order of lipid phases in model and plasma membranes, Proc. Natl Acad. Sci. USA 106 (2009) 16645–16650. [30] D.T. Browman, M.B. Hoegg, S.M. Robbins, The SPFH domain-containing proteins: more than lipid raft markers, Trends Cell Biol. 17 (2007) 394–402. [31] I.C. Morrow, R.G. Parton, Flotillins and the PHB domain protein family: rafts, worms and anaesthetics, Traffic 6 (2005) 725–740. [32] E. Rivera-Milla, C.A.O. Stuermer, E. Malaga-Trillo, Ancient origin of reggie (flotillin) and reggie-like proteins and convergent evolution of SPFH domain, Cell. Mol. Life Sci. 63 (2006) 343–357. [33] G.P. Solis, M. Hoegg, C. Munderloh, E. Rivera-Milla, C.A.O. Stuermer, Rules for reggie oligomerization and stabilization, Biochem. J. 403 (2007) 313–322. [34] K. Simons, R. Ehehalt, Cholesterol, lipid rafts, and disease, J. Clin. Invest. 110 (2002) 597–603. [35] K.G. Rothberg, J.E. Heuser, W.C. Donzell, Y.S. Ying, J.R. Glemney, R.G. Anderson, Caveolin, a protein component of caveolae membrane coats, Cell 68 (1992) 673–682. [36] A.M. Fra, E. Williamson, K. Simons, R.G. Parton, De novo formation of caveolae in lymphocytes by expression of VIP21–caveolin, Proc. Natl Acad. Sci. USA 92 (1995) 8655–8659. [37] M.F. Langhorst, A. Reuter, C.A.O. Stuermer, Scaffolding microdomains and beyond —the function of reggie/flotillin proteins, Mol. Cell. Life Sci. 62 (2005) 2228–2240. [38] M. Kirkham, S.J. Nixon, M.T. Howes, L. Abi-Rached, D.E. Wakeham, M. HanzalBayer, C. Ferguson, M.M. Hill, M. Fernandez-Rojo, D.A. Brown, J.F. Hancock, F.M. Brodsky, R.G. Parton, Evolutionary analysis and molecular dissection of caveola biogenesis, J. Cell Sci. 121 (2008) 2075–2086. [39] M. Drab, P. Verkade, M. Elger, M. Kasper, M. Lohn, B. Lauterbach, J. Menne, C. Lindschau, F. Mende, F.C. Luft, Loss of caveolae, vascular dysfunction, and pulmonary defects in caveolin-1 gene-disrupted mice, Science 293 (2001) 2449–2452. [40] S.J. Nixon, A. Carter, J. Wegner, C. Ferguson, M. Floetenmeyer, J. Riches, B. Key, M. Westerfield, R.G. Parton, Caveolin-1 is required for lateral line neuromast and notochord development, J. Cell Sci. 120 (2007) 2151–2161. [41] P.-K. Fang, K.R. Solomon, L. Zhuang, Q. Maosong, M. McKee, M.R. Freeman, P.C. Yelick, Caveolin-1alpha and -1beta perform nonredundant roles in early vertebrate development, Am. J. Pathol. 169 (2006) 2209–2222. [42] M. Murata, J. Peranen, R. Schreiner, F. Wieland, T.V. Kurzchalia, K. Simons, VIP21/ caveolin is a cholesterol-binding protein, Proc. Natl Acad. Sci. USA 92 (1995) 10339–10343. [43] J.L. Thorpe, M. Doitsidou, S.-Y. Ho, E. Raz, S.A. Farber, Germ cell migration in zebrafish is dependent on HMGCoA reductase activity and prenylation, Dev. Cell 6 (2004) 295–302. [44] S.J. Nixon, J. Wegner, C. Ferguson, P.-F. Méry, J.F. Hancock, P.D. Currie, B. Key, M. Westerfield, R.G. Parton, Zebrafish as a model for caveolin-associated muscle 422 [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] C.A.O. Stuermer / Biochimica et Biophysica Acta 1812 (2011) 415–422 disease; caveolin-3 is required for myofibril organization and muscle cell patterning, Hum. Mol. Genet. 14 (2005) 1727–1743. Z. Tang, P.E. Scherer, T. Okamoto, K. Song, C. Chu, D.S. Kohtz, I. Nishimoto, H.F. Lodish, M.P. Lisanti, Molecular cloning of caveolin-3, a novel member of the caveolin gene family expressed predominantly in muscle, J. Biol. Chem. 271 (1996) 2255–2261. T. Benzing, Signaling at the slit diaphragm, J. Am. Soc. Nephrol. 15 (2004) 1382–1391. S. Roselli, O. Gribouval, N. Boute, M. Sich, F. Benessy, T. Attie, M.C. Gubler, C. Antignac, Podocin localizes in the kidney to the slit diaphragm area, Am. J. Pathol. 160 (2002) 131–139. T.B. Huber, M. Simons, B. Hartleben, L. Sernetz, M. Schmidts, E. Gundlach, M.A. Saleem, G. Walz, T. Benzing, Molecular basis of the functional podocin–nephrin complex: mutations in the NPHS2 gene disrupt nephrin targeting to lipid raft microdomains, Hum. Mol. Genet. 12 (2004) 3397–3405. A.G. Kramer-Zucker, S. Wiessner, A.M. Jensen, I.A. Drummond, Organization of the pronephric filtration apparatus in zebrafish requires Nephrin, Podocin and the FERM domain protein Mosaic eyes, Dev. Biol. 285 (2005) 316–329. I.A. Drummond, Kidney development and disease in the zebrafish, Am. Soc. Nephrol. 16 (2005) 299–304. S. Roselli, I. Moutkine, O. Gribouval, A. Benmerah, C. Antignac, Plasma membrane targeting of podocin through the classical exocytic pathway: effect of NPHS2 mutations, Traffic 5 (2004) 37–44. S.O. Deininger, L. Rajendran, F. Lottspeich, M. Przybylski, H. llges, C.A.O. Stuermer, A. Reuter, Identification of teleost Thy-1 and association with the microdomain/ lipid raft reggie proteins in regenerating CNS axons, Mol. Cell. Neurosci. 22 (2003) 544–554. A. Reuter, E. Malaga-Trillo, U. Binkle, E. Rivera-Milla, R. Beltre, Y. Zhou, M. Bastmeyer, C.A.O. Stuermer, Evolutionary analysis and expression of teleost Thy1, Zebrafish 1 (2004) 191–201. E. Malaga-Trillo, G.P. Solis, V. Tomanetz, C. Geis, L. Luncz, C.A.O. Stuermer, Regulation of embryonic cell adhesion by the prion protein, PLoS Biol. 7 (2009) e1000055. C.A. Baumann, V. Ribon, M. Kanzaki, D.C. Thurmond, S. Mora, S. Shigematsu, P.E. Bickel, J.E. Pessing, A.R. Saltiel, CAP defines a second signalling pathway required for insulin-stimulated glucose transport, Nature 407 (2000) 202–207. [56] N. Kioka, K. Ueda, T. Amachi, Vinexin, CAP/ponsin, ArgBP2: a novel adaptor protein family regulating cytoskeletal organization and signal transduction, Cell Struct. Funct. 27 (2002) 1–7. [57] V.N.B. Das, A. Dujeancourt, M.I. Toulouze, T. Galli, P. Roux, A. Dautry-Varsat, A. Alcover, Activation-induced polarized recycling targets T cell antigen receptors to the immunological synapse: involvement of the SNARE complexes, Immunity 20 (2004) 577–588. [58] L. Chang, S.-H. Chiang, A.R. Saltiel, TC10alpha is required for insulin-stimulated glucose uptake in adipocytes, Endocrinology 148 (2007) 27–33. [59] L.A. Feig, Ral-GTPases: approaching their 15 minutes of fame, Trends Cell Biol. 13 (2003) 419–425. [60] G. Lalli, A. Hall, Ral GTPases regulate neurite branching through GAP-43 and the exocyst complex, J. Cell Biol. 171 (2005) 657–669. [61] M. Langhorst, A. Reuter, M. Wiechers, S. Hannbeck, H. Plattner, C.A.O. Stuermer, Trafficking and association with PrPc of the microdomain proteins reggie-1 and reggie-2 (flotillins), Eur. J. Cell Biol. 87 (2008) 211–226. [62] M.F. Langhorst, C. Munderloh, S. Mueller, F. Jaeger, S. Hartmann, G. Luxenhofer, M. Wiechers, C.A.O. Stuermer, Reggies/flotillins regulate Rho GTPase signalling during axon outgrowth and axon regeneration, Eur. J. Cell Biol. 87 (2008) 921–931. [63] X.-W. Chen, M. Inoue, S.C. Hsu, A.R. Saltiel, RalA-exocyst-dependent recycling endosome trafficking is required for the completion of cytokinesis, J. Biol. Chem. 281 (2006) 38609–38616. [64] M.F. Langhorst, S. Hannbeck von Hanwehr, G.P. Solis, H. Plattner, C.A.O. Stuermer, Linking membrane microdomains to the cytoskeleton: regulation of the lateral mobility of reggie-1/flotillin-2 by interaction with actin, FEBS Lett. 581 (2007) 4697–4703. [65] J. Liu, S.M. DeYoung, M. Zhang, L.H.S.A.R. Dold, The stomatin/prohibitin/flotillin/ HflK/C domain of flotillin-1 contains distinct sequences that direct plasma membrane localization and protein interactions in 3T3-L1 adipocytes, J. Biol. Chem. 280 (2005) 16125–16134. [66] C.G. Becker, B.C. Lieberoth, F. Morellini, J. Feldner, T. Becker, M. Schachner, L1.1 is involved in spinal cord regeneration in adult zebrafish, J. Neurosci. 24 (2004) 7837–7842.
© Copyright 2025 Paperzz