Molecular Vision 2013; 19:1779-1794 <http://www.molvis.org/molvis/v19/1779>
Received 27 May 2013 | Accepted 02 August 2013 | Published 06 August 2013

Identification of HMX1 target genes: A predictive promoter model approach

Arnaud Boulling,1 Linda Wicht,1,2 Daniel F. Schorderet1,2,3

1Institute for Research in Ophthalmology, Sion, Switzerland; 2School of Life Sciences, Federal Institute of Technology (EPFL), Lausanne, Switzerland; 3Department of Ophthalmology, University of Lausanne, Lausanne, Switzerland

Correspondence to: Daniel F. Schorderet, IRO, Av. du Grand-Champsec 64, 1950 Sion, Switzerland; Phone: +4127 205 7900; FAX: +4127 205 7901; email: daniel.schorderet@irovision.ch

Abstract

Purpose: A homozygous mutation in the H6 family homeobox 1 (HMX1) gene is responsible for a new oculoauricular defect leading to eye and auricular developmental abnormalities as well as early retinal degeneration (MIM 612109). However, the HMX1 pathway remains poorly understood, and in the first approach to better understand the pathway’s function, we sought to identify the target genes.

Methods: We developed a predictive promoter model (PPM) approach using a comparative transcriptomic analysis in the retina at P15 of a mouse model lacking functional Hmx1 (dmbo mouse) and its respective wild-type. This PPM was based on the hypothesis that HMX1 binding site (HMX1-BS) clusters should be more represented in promoters of HMX1 target genes. The most differentially expressed genes in the microarray experiment that contained HMX1-BS clusters were used to generate the PPM, which was then statistically validated. Finally, we developed two genome-wide target prediction methods: one that focused on conserving PPM features in human and mouse and one that was based on the co-occurrence of HMX1-BS pairs fitting the PPM, in human or in mouse, independently.

Results: The PPM construction revealed that sarcoglycan, gamma (35kDa dystrophin-associated glycoprotein) (Sgcg), teashirt zinc finger homeobox 2 (Tshz2), and solute carrier family 6 (neurotransmitter transporter, glycine) (Slc6a9) genes represented Hmx1 targets in the mouse retina at P15. Moreover, the genome-wide target prediction revealed that mouse genes belonging to the retinal axon guidance pathway were targeted by Hmx1. Expression of these three genes was experimentally validated using a quantitative reverse transcription PCR approach. The inhibitory activity of Hmx1 on Sgcg, as well as protein tyrosine phosphatase, receptor type, O (Ptpro) and Sema3f, two targets identified by the PPM, were validated with luciferase assay.

Conclusions: Gene expression analysis between wild-type and dmbo mice allowed us to develop a PPM that identified the first target genes of Hmx1.

Introduction

The homeobox (HMX) family of transcription factors is characterized by the presence of a 60-amino acid homeobox domain. Currently, this family contains four members: HMX1, HMX2, HMX3 (also known as Nkx5–3, Nkx5–2, and Nkx5–1, respectively), and sensory organ homeobox 1 (SOHo1) [1]. Expression of HMX1, HMX2, and HMX3 is highest in the sensory organs, I.E., the eye and inner ear, and in the peripheral and central nervous systems [2,3]. During mouse development, Hmx1 is expressed in the trigeminal ganglion and in the second branchial arches early as E9.5, and in the dorsal root ganglia at E10.5. Later on,Hmx1 is expressed in the lens, in the neural epithelium of the eye, in the sympathic and vagal nerve ganglia, and in the mesenchyme near the developing ear [4]. More recently, the discovery of an HMX1 loss-of-function mutation responsible for a new oculoauricular syndrome (MIM 612109) in a Swiss consanguineous family prompted us to evaluate the role of this transcription factor [5].

In 2009, the description of two mutant mice called “dmbo” and “misplaced ears” exhibiting microphthalmia, in addition to ear and cranial malformations, was reported. Mapping and sequencing analyses of these mice revealed a nonsense mutation in the first exon of Hmx1 in dmbo and a frameshift mutation in exon 2 of “misplaced ears” mice [6]. The absence of Hmx1 protein in dmbo was further confirmed in a study showing that Hmx1 was required for the normal development of somatosensory neurons in the geniculate ganglion [7]. Moreover, a dmbo rat strain with a similar phenotype and a deletion in an ancient distal putative enhancer of Hmx1 was described recently [8]. All these rodent mutants underline the prominent role of Hmx1 in eye development and represent good models.

Despite these recent advances, the role of Hmx1 in transcriptional regulation remains widely unknown. A major challenge in deciphering the Hmx1 pathway involved in eye development is identifying target genes. However, this represents a difficult task as no HMX1 chromatin immunoprecipitation-grade antibody seems to exist in the mouse. Therefore, we constructed a predictive promoter model (PPM). This approach is based on the analysis of differentially co-expressed genes between two different biologic states and represents a powerful tool as was recently shown [9]. In our case, we used a comparative transcriptomic analysis between retinas at postnatal day 15 (P15) of wild-type (WT) and dmbo mice. Basically, a promoter model is defined as a framework of two or more transcription factor binding sites (TFBSs) with a defined distance range and strand orientation. In a given promoter, a functional pattern involving multiple TFBSs is called a cis-regulatory module (CRM). CRMs represent the next level of organization after individual TFBSs and are often involved in tissue-specific expression (reviewed in [10] and [11]). The promoter model is called predictive when it is based on a functional hypothesis instead of the analysis of experimentally validated TFBSs. In theory, the promoter model represents a specific and flexible structure shared by the promoter of genes belonging to the same pathways.

Despite the lack of knowledge about Hmx1, critical information was sufficient to identify specific features about the gene’s target. In fact, Amendt et al. showed that HMX1 binds to the canonical CAAGTG sequence and acts as a transcriptional antagonist of Nkx2-5, a well-studied transcription factor that recognizes a consensus sequence TNAAGTG overlapping HMX1-BSs [12]. The mouse and rat ANF proximal promoters include two validated Nkx2-5-BSs involved in transcriptional activation. Additional sites are located in distal enhancer regions upstream of the transcription start site (TSS) [13-15]. Similar Nkx2–5-BSs clusters have also been observed in the H15 mid locus of Drosophila. The functionality of one of them has been demonstrated in cardioblasts [16]. This type of CRM involving multiple similar TFBSs is called homotypic CRMs, or homotypic clusters of TFBSs, and is widely represented in proximal promoters and enhancers of mammals and invertebrates [17]. This observation is particularly true for TFBSs of several TFs, including Nkx2–5.

To identify targets of HMX1, we developed a PPM based on HMX1-BSs clusters, with analogy to Nkx2–5. We report the first Hmx1 targets in the mouse retina at P15. Moreover, applying our PPM to mouse and human genomes allowed us to identify additional potential target genes involved in embryonic eye development.

Methods

Animal handling and tissue isolation

The studies adhered to the Association for Research in Vision and Ophthalmology (ARVO) Statement for the Use of Animals in Ophthalmic and Vision Research and were approved by the Veterinary Service of the State of Valais (Switzerland). WT C57BL/6J mice were obtained from Janvier (Le Genest St Isle, France) and dmbo mice from Jackson Laboratory (Bar Harbor, ME). Dmbo mice were backcrossed with C57BL/6JWT mice for three additional generations to obtain a homogeneous genetic background and to remove the Rd1 mutation they unexpectedly carried. All mice were genotyped with polymerase chain reaction analysis of DNA tails. Animals were maintained in a 12 h:12 h light-dark cycle with free access to food and water. Mice were killed by cervical dislocation at P15 or P60, and their eyes were enucleated. Retinas were isolated under a microscope to remove extra retinal tissue and snap-frozen at −80 °C.

RNA extraction and dosage

Total RNA was individually isolated from each whole retina and purified using the RNeasy minikit (Qiagen, Basel, Switzerland) as described by the manufacturer. RNA quantities were assessed with a NanoDrop ND-1000 spectrophotometer (Thermo Scientific, Wilmington, DE). Four and three different animals for each condition were used for microarray and reverse transcription polymerase chain reaction (RT–PCR), respectively.

Microarray procedure

RNA quality was assessed using RNA 6000 NanoChips with the Agilent 2100 Bioanalyzer (Agilent, Palo Alto, CA). For each sample, 100 ng of total RNA were amplified using the WT sense strand Target Labeling kit (Affymetrix, Santa Clara, CA); 5.5 µg of the resulting sense cDNA was fragmented with uracil DNA glycosylase (UDG), apurinic/apyrimidic endonuclease 1 (APE 1), and biotin-labeled with terminal deoxynucleotidyltransferase (TdT) using the GeneChip WT Terminal labeling kit (Affymetrix). Affymetrix Mouse Gene 1.0 ST arrays were hybridized with 2.7 µg of biotinylated target for 17 h at 45 °C, washed, and stained according to the protocol described in the Affymetrix GeneChip Expression Analysis Manual (Fluidics protocol FS450_0007). The arrays were scanned using the GeneChip Scanner 3000 7G (Affymetrix), and raw data were extracted from the scanned images and analyzed with the Affymetrix Power Tools software package. All statistical analyses were performed using the free high-level interpreted statistical language R (version 2.12.1) and the Bioconductor package limma (version 3.6.9). Hybridization quality was assessed using the Expression Console software (Affymetrix). Normalized expression signals were calculated from Affymetrix CEL files using the RMA normalization method. Differential hybridized features were identified using Bioconductor package “limma” that implements linear models for microarray data [18]. The p values were adjusted for multiple testing with Benjamini and Hochberg’s method to control for the false discovery rate (FDR) [19]. Probe sets showing at least 1.2-fold change and a FDR<0.1 were considered significant. Gene expression data have been deposited in GEO (GSE47002).

Functional annotation of microarray data

Differentially expressed genes (FDR<0.1) were annotated in accordance with the Gene Ontology (GO) classification system. GO terms were classified into categories related to molecular function, cell component, and biologic process to assess the statistical enrichment of differentially expressed genes in these categories compared with the full mouse genome. Annotation and statistical calculation were realized using the DAVID algorithm [20,21]. In addition, the MetaCore software from GeneGo was used to highlight the most relevant GeneGO process networks. Each process represents a preset network of protein interactions characteristic of the process. For the DAVID and MetaCore enrichment analyses, only results with a p value <0.1 were considered. MetaCore and DAVID use a hypergeometric model to determine the significance of the enrichment.

Predictive promoter model construction and validation

All sequences were collected via the UCSC Main Table Browser of the online Galaxy Platform (https://main.g2.bx.psu.edu/). We used the July 2007 (NCBI37/mm9) and February 2009 (GRCh37/hg19) genome assemblies’ versions, and genes absent in the refGene table were retrieved in GenBank and checked manually. All gene accession numbers related to the genes cited in the article are summarized in Appendix 1. [-250;+250] region selection and motif combinations analyses in the (+) and (-)-training sets were obtained from Galaxy. The (-)-training set was constituted by random selection of 2,000 RefSeq gene promoter sequences. Finally, statistical validation of the model was performed with Fisher’s exact test. PPM specificity and sensitivity were calculated with MedCalc. Cell-specific expression levels of Hmx1, sarcoglycan, gamma (35kDa dystrophin-associated glycoprotein) (Sgcg), teashirt zinc finger homeobox 2 (Tshz2), and solute carrier family 6 (neurotransmitter transporter, glycine) (Slc6a9) were retrieved in the gene expression profile database [22].

Predictive promoter model–based genome-wide target predictions

All the [-250,+200] sequences fitting the PPM were selected from human and mouse RefSeq databases. PPM-based selection focused on the conserved HMX1-BS pairs was realized by crossing the human and mouse previously obtained selections. Sequences containing more than two HMX1-BSs were analyzed to achieve the target genes selection based on the co-occurrence of HMX1-BS pairs. Axon guidance pathway enrichment analysis was based on the KEGG database and performed for the mouse predicted target selection obtained with the PPM co-occurrence based method. The statistical enrichment was assessed against the full mouse genome by considering the 25,504 unique genes of the ccds table, with the χ2 and Yates' correction test.

Quantitative rt-PCR

Reverse transcription was performed with 500 ng RNA, 25 ng/µl OligodT, 1 mM each dNTP, 10 mM dithiothreitol, 2 U/µl RNaseBlock, and 1 µl AffinityScript (Agilent) in a final volume of 20 µl at 42 °C for 1 h. The reaction was then maintained at 70 °C for 15 min, and the cDNA obtained was 1:10 diluted. Quantitative PCR was performed in a 25-µl mixture containing 12.5 µl of FastStart Universal SYBR Green Master (ROX; Roche, Basel, Switzerland), 10 µl of diluted cDNA, and 0.3 µM of primer pairs (Appendix 2). The PCR program had an initial denaturation at 95 °C for 10 min, followed by 40 cycles of denaturation at 95 °C for 30 s, annealing at 55 °C for 1 min, and extension at 72 °C for 1 min. All PCRs were realized in triplicate. Transcript levels were normalized using the Gapdh housekeeping gene and analyzed with the Student t test. All qPCR efficiencies were calculated from the slope of a standard dilution curve to allow relative comparison of gene mRNA levels.

Immunohistochemistry

Enucleated eyes were fixed for 45 min at 4 °C in 4% paraformaldehyde and were cryoprotected by 30% sucrose. The eyes were then embedded in freezing compound (30% Albumin/3% gelatin in 1x phosphate-buffered saline (PBS): 154 mM NaCl, 1 mM KH2PO4, 3 mM Na2HPO4 heptahydrate) and vertically sliced 10 μm thick in a cryostat. Sections were washed three times with PBS, treated with blocking buffer (2% native goat serum containing 0.2% Triton X-100) at room temperature for 10 min, and left overnight at 4 °C with the anti-SGCG primary antibody (Proteintech, Chicago, IL) diluted 1:100. Controls were prepared by omitting the primary antibody during the incubation. The following morning, sections were rinsed three times with PBS, blocked for 10 min, and incubated at room temperature for 1 h with Alexa Fluor 594 goat antirabbit immunoglobulin (Invitrogen, Zug, Switzerland) secondary antibody diluted 1:1,500. After three additional washings, the sections were stained with 4',6-diamidino-2-phenylindole dihydrochloride diluted 1:1,500 for 10 min at room temperature, washed three times again, and mounted with Citifluor AF1 (Citifluor Ltd, Leicester, UK). The stained slides were imaged on a Zeiss microscope, and image analysis was performed using the ZEN lite 2011 software (Zeiss, Zürich, Switzerland).

Construction of reporter and expression plasmids

PCR primer pairs used for molecular cloning were designed to generate an amplicon spanning the TSS of Sgcg, Sema3f, and Ptpro and to carry all the HMX1-BS identified in the proximal promoter region. All promoter regions were amplified with the PfuUltra High-Fidelity DNA Polymerase (Agilent) according to the manufacturer’s instructions. Mouse Sgcg, Sema3f, and Ptpro promoter regions were amplified using the following primer pairs carrying MluI and XhoI restriction sites (underlined), with indicated annealing temperatures (Ta): 5′-GCG C AC GCG TCA AAG ACA CGT CAG CCT CAG-3′ and 5′-GCG C CT CGA GGA AAC GCT GTA CCT ATC TGA TTT ACA-3′ (Sgcg, Ta=61 °C), 5′-GCG C AC GCG TGC AAG AGT GTA TGG GGA AGG-3′ and 5′-GCG C CT CGA GCA GGC CTC TCA GCA GGTG-3′ (Sema3f, Ta=63 °C), 5′-GCG C AC GCG TCA TGG AAA TCG TTG CTT GTG-3′ and 5′-GCG C CT CGA GCG GCG TTG TTT AAT GGC TAA-3′ (Ptpro, Ta=60 °C). Amplified DNA fragments were inserted into the pGL3-basic vector with the XhoI and MluI restriction enzymes to produce pGL3-Sgcg, pGL3-Sema3f, and pGL3-Ptpro reporter constructs. Before the cloning step, the CAAGTG site located just upstream of the multiple cloning site in the pGL3-basic vector was converted to TAATCA by site-directed mutagenesis. The Hmx1 mouse cDNA was amplified with PfuUltra High-Fidelity DNA Polymerase using 5′-ATG CCG GAT GAG CTG ACC G-3′ and 5′-TCA CAC TAG CCC CGG CAT C-3′ primers (Ta=60 °C), and then inserted into the pcDNA3.1 vector (pcDNA3.1-Hmx1) with the pcDNA3.1/V5-His TOPO TA Expression Kit (Invitrogen).

Cell culture and transfection

Mouse neuroblastoma cells (aka Neuro-2a or N2a) were grown in Dulbecco’s Modified Eagle Medium (DMEM; PAA, Cölbe, Germany) supplemented with 10% fetal bovine serum and 100 µg/ml Normocin (Invivogen, Toulouse, France). Twenty-four hours before transfection, 200,000 cells/well were seeded in 12-well plates. For one transfection, 900 ng of one pGL3 reporter construct, 900 ng of pcDNA3.1-Hmx1 or empty pcDNA3.1, and 300 ng of pCMV-Beta-Gal control plasmid were mixed together with 4 µl jetPEI (Polyplus, Illkirch, France) and dropped onto the cells.

Luciferase reporter gene assay

Forty-eight hours after transfection, cells were rinsed with PBS and lysed with 100 µl potassium phosphate buffer (100 mM K2HPO4, pH7.8, 0.2% Triton X-100). After centrifugation, 5 µl supernatant from each sample were added to 20 µl Firefly luciferase reagent. In parallel, 10 µl supernatant from each sample were added to 100 µl β-galactosidase reagent. The relative luciferase activity was determined by dividing the luminescence of Firefly luciferase activity by that of the cotransfected β-galactosidase activity. The experiment was performed three times for each reporter construct, and transfections were realized in triplicate for each experiment. The significance between the luciferase activity of each reporter construct cotransfected with pcDNA3.1-Hmx1 and with pcDNA3.1 was then assessed for significance with the Student t test.

Statistical analysis

The statistical tests used in this study are detailed at the end of the microarray procedure, functional annotation of microarray data, PPM construction and validation, quantitative reverse transcription PCR, and luciferase reporter gene assay sections.

Results

Comparative transcriptomic analysis

The comparative transcriptomic analysis of the mouse retina between the dmbo and WT C57BL/6J mice was realized at P15 to avoid killing pregnant dmbo mice and to obtain a sufficient amount of tissue. The retina is still developing at this age and is always expressing Hmx1. The analysis showed 146 differentially expressed genes (70 up and 76 down) with a FDR<0.1 and at least 1.2-fold change (Figure 1A, Appendix 3, Appendix 4). Thirty of these genes were highly confident and had a FDR<0.01 (14 up and 16 down). Analysis of the 146 differentially expressed genes with the MetaCore software from GeneGo showed ten enriched GeneGO process networks with p<0.1 (Figure 1B). The first three ranked processes are the synaptogenesis (p=0.008879), the visual perception (p=0.009481), and the synaptic contact (p=0.009713). In another approach, the DAVID software allowed us to classify these genes into GO categories, showing that nine of them were significantly enriched with p<0.1 in molecular function, three in cell component, and 20 in biologic process categories (Appendix 5). All of the enriched molecular function categories were related to ion and vitamin binding or transmembrane transport, and all the enriched cell component categories were related to cell projection terms as axoneme and cilium. Enriched biologic process categories were more numerous and diversified and concerned, for example, organelle localization or metal ion homeostasis.

Hmx1 target promoter model construction and validation

As explained above, we hypothesized that multiple Hmx1 sites could form CRMs and act in synergy, as observed for Nkx2–5 [17]. In this regard, we postulated that the number of motifs could play a role in the transcriptional regulation of Hmx1 targets. To elaborate our predictive promoter model, we used the promoter sequences of the most confident differentially expressed genes (FDR<0.01 group) and considered the number of CAAGTG motifs present. We used a screening window ranging in size from −250 to +200 nucleotides (nt) around the TSS. This window was based on the size of the proximal promoter and the approximate median size of the eukaryote 5′ untranslated region (UTR), two regions known to be enriched in TFBS [23,24]. The CAAGTG motifs located in the [-250,+200] interval were counted. Three of the 30 genes contained two motifs, three contained one motif, and the last 24 did not contain any motif. The three genes containing two motifs, considered theoretical Hmx1 targets, were used to generate a framework (Figure 2A–B) that should correspond to a feature specific for Hmx1 target promoters. The orientation of the motifs was disregarded, but we considered the space between the two motifs and kept a distance range spanning from 90 to 190 nt. Unexpectedly, only one HMX1-BS in the Tshz2 promoter was strictly conserved between human and mouse.

This theoretical promoter model is predictive and must be statistically validated, as it is based on arbitrary criteria. Therefore, we calculated the enrichment of this particular feature between the FDR<0.01 group (used as the positive (+)-training set) and a control group of random promoters (used as the negative (-)-training set). We observed a 22.2-fold significant enrichment (p=0.0006) of the PPM specific feature in the (+)-training set compared to the (-)-training set within the [-250,+200] region (Figure 2C). This enrichment was also observed in the FDR<0.1 group but to a lesser extent (p=0.0091). There was no significant enrichment when the genes containing at least one motif were considered. From these data, we concluded that the high enrichment of this particular feature in the (+)-training set validated our promoter model and supported the idea that Sgcg, Tshz2, and Slc6a9 were direct Hmx1 targets.

To complete our approach, we also tried to build a PPM with lower specificity and higher sensitivity. For this, we replaced the canonical sequence CAAGTG with the minimal core motif CAAG also able to bind HMX1 but with a lower affinity [12]. We generated a low specific PPM (LS-PPM) and a very low specific PPM (VLS-PPM) fitting the same distance and orientation criteria than the initial PPM, but carrying one or two CAAG in place of the canonical HMX1-BS, respectively (Table 1). Both retrieved a better rate of positive genes but showed low specificity (1.37- and 1.01-fold enrichment for LS-PPM and VLS-PPM, respectively). These two PPMs with lower specificity were not reused for the following analyses.

Characterization of Sgcg, Tshz2, and, Slc6a9 expression

To minimize the possibility that Sgcg, Tshz2, and Slc6a9 were false positive targets of Hmx1, we confirmed their level of deregulation between eyes from dmbo and WT mice, and checked for colocalization with Hmx1. The retina is made of many different cell types with specific expression profiles. We therefore assessed the expression of Hmx1, Sgcg, Tshz2, and Slc6a9 with qPCR (Figure 3A) and looked for their precise cell subtype expression in the mouse retina, according to an online gene expression profile database [22] (Figure 3B). Transcript quantification confirmed that Sgcg and Tshz2 were overexpressed in dmbo at P15, whereas Slc6a9 was underexpressed. This deregulation tended to disappear at P60 for Tshz2 and Slc6a9, but remained extremely high for Sgcg. As expected, the inspection of a retina-specific database revealed a strong overlap of Hmx1, Slc6a9, and Tshz2 expression in the glycinergic amacrine cells. Sgcg expression was not detected in the microarray database probably due to the weak level of mRNA, as shown with qPCR. However, the γ-sarcoglycan protein encoded by Sgcg was detected with immunohistochemistry in the ganglion cell (GCL), the inner plexiform (IPL), the inner nuclear (INL), and the outer plexiform layers (OPL; Figure 3C), as already shown by Fort et al. [25]. No difference in protein expression of γ-sarcoglycan was observed between the WT and dmbo samples (data not shown). The immunohistochemistry staining was higher in some cells of the INL exhibiting a disposition pattern characteristic of the amacrine cells, at the delimiting border between the INL and the IPL. This result was in accordance with colocalization of Sgcg and Hmx1 expression.

Predictive promoter model–based genome-wide screening for HMX1 putative targets

Our model was based on a comparative transcriptomic analysis realized in the mouse retina at P15. However, Hmx1 is highly expressed in the mouse eye as early as E10.5 suggesting an important role in development. Assuming that the PPM we developed was specific for HMX1 targets (with 0.45% versus 10% representation in the (-) and (+)-training sets, respectively), we decided to screen the full human and mouse genomes. This global approach should provide an exhaustive view of all putative HMX1 targets, including those expressed during embryonic eye development. As the first step, we used the PPM to screen mouse and human RefSeq databases via the Galaxy platform. These two databases contained a total of 30,490 and 43,695 sequences respectively, which corresponded to all transcripts of reference, including all isoforms and alternative promoters. We considered for each gene all potential alternative promoters whereas all redundant promoter sequences due to alternative splicing were discarded. The gene accession numbers of all isoforms corresponding to our predicted genes are listed in Appendix 1.

Screening of the full mouse RefSeq database using the PPM within the [-250,+200] region retrieved 157 sequences corresponding to unique protein-coding genes (Figure 4A, Appendix 1). Similarly, the screening of the human genome retrieved 100 sequences corresponding to unique protein-coding genes. This approach allowed us to generate an exhaustive list of all possible HMX1 targets, but the high number of positive hits resulted in some difficulties with their interpretation. In fact, some of these genes were probably true HMX1 targets, but several could also be false-positives, representing targets related to TFs with the same binding sites (for example, HMX3 or NKX2–5). To improve the selectivity of the analysis and to focus on the most interesting candidate genes, we decided to add additional filters.

Initially, we developed a PPM approach based on conserving the HMX1-BS pairs between human and mouse. In fact, comparative genomics is one of the usual methods that aimed at discriminating functional TFBSs from irrelevant ones (reviewed in [11]). To maintain relative flexibility, our method was based only on the presence of an HMX1-BS pair and did not implicate a strict conservation of the positions or orientations of the HMX1-BSs between the two species. As already demonstrated, traditional approaches, similar to phylogenetic footprinting, give good predictions but are also likely to miss important conserved regulatory elements [26]. Crossing the two data sets showed that only ten genes contained the PPM features in both species (Figure 4B). These genes were classified according to the localization of their expression.

In a second phase, we used another PPM approach based on the cooccurrence of HMX1-BS pairs, driven by the basic idea that increasing the number of HMX1-BS in the promoter region should increase their interaction with HMX1. This phenomenon should result in more efficient transcription regulation. To assess this hypothesis, we looked at the HMX1-BS occurrence in all the mouse and human promoter sequences fitting the PPM. We observed that the promoter regions contained a maximum of four HMX1-BSs within the [-250,+200] window. Always in accordance with the PPM criteria, we determined that three sites might form three different homotypic HMX1-BS pairs (P1–2,P1–3,P2–3), even if it is unlikely that all three pairs could be considered at the same time due to the minimum distance range constraint of 90 bp (Figure 5A). Similarly, four sites might lead to a maximum of five combinations (P1–2, P1–3,P2–3, P2–4, P3–4). A single given site can be involved in multiple combinations. The screening of the mouse genome with this method retrieved nine genes with three sites allowing two different pair combinations, and one gene (Ephrin type-A receptor 6; Epha6) with four sites allowing three different pair combinations (Figure 5B). Interestingly, three of these ten genes (Epha6, Ptpro, and Sema3f) are expressed in the retina and are involved in axonal growth repulsion (see discussion). It represents a 56.6-fold enrichment (p<0.0001) in the mouse axon guidance KEGG pathway (mmu04360). Ptpro was incorporated within the pathway although this gene was not initially reported in the KEGG database, in spite of Ptpro’s role as a guidance cue in retinal neurons [27,28]. Moreover, a deeper examination of Epha6 showed an additional HMX1-BS at position [+245,+250] and an additional HMX1-BSs cluster fitting the PPM within the first intron (Figure 5C). Among the mouse HMX1-BSs, two were conserved in the human EPHA6 proximal promoter. One was unique to the human gene. With a similar approach, we identified eight genes with three sites allowing two interactions in the human genome (Figure 5B).

Experimental validation of several predicted targets with luciferase assay

We experimentally validated some of these results with luciferase assays performed in N2A cells. We first assessed the reliability of our system by cotransfecting the pGL3-Sgcg positive control reporter construct with the pcDNA3.1-Hmx1 expression construct. Hmx1 cotransfection decreased pGL3-Sgcg luciferase expression by 71%, which was expected given the dmbo qPCR Sgcg results (Figure 6). Cotransfection of Hmx1 repressed pGL3-Ptpro and pGL3-Sema3f luciferase expression by 50% and 66%, respectively.

Discussion

The major goal of this study was to identify target genes of Hmx1 in the mouse retina. The integration of in vitro data from Ament et al. and our in vivo microarray data allowed us to obtain a clear picture of the typical basic structure of an HMX1 target promoter [12]. In addition, experimental controls concerning transcript amounts, expression colocalization, and luciferase assay strongly supported these findings.

The microarray data yielded the first set of information about the molecular phenotype of the dmbo retina. The MetaCore analysis underlined that a lack of HMX1 protein altered synaptogenesis and visual perception, two biologic processes occurring at P15 [29]. At this time, we cannot say whether these observations were directly linked to Hmx1 loss of activity or if they derived from anterior impairments occurring during eye development.

Then, we used the microarray data to construct a high specific PPM based on HMX1-BS clusters. It revealed that Sgcg, Tshz2, and Slc6a9 were Hmx1 targets in the mouse retina at P15. Using degenerated binding motifs for PPM construction, such as the minimal CAAG core motif, led to higher sensitivity but also to low specificity with no significant enrichment. Such a low specific PPM cannot be used for prediction because the model would probably yield an extremely high number of false positives. However, we thought that Hmx1 probably binds degenerate motifs in vivo, but no position weight matrix is currently available to perform a better PPM for Hmx1. Moreover, the sensitivity of the original PPM is likely underestimated because some of the differential expressions observed for genes belonging to the (+)-training set probably result from secondary events and are not directly linked to Hmx1.

The expression of Hmx1, Tshz2, Slc6a9, and probably Sgcg was observed in the glycinergic amacrine cells, a cell type that establishes synaptic contacts with rod-driven bipolar cells and play an important role in neurotransmission. Tshz2 is involved in an axonal growth network in the mouse retina and is expressed in the zebrafish neural retina at 48 h post fertilization [30,31]. Slc6a9 is specifically expressed in the glycinergic amacrine cells where it plays an important role in glycine uptake, and controls N-methyl-D-aspartic acid receptor coagonist occupancy in the mouse retina [32]. The role of both genes in the retina needs to be further investigated. However, their expression seemed to be totally or partially compensated at P60, suggesting that Hmx1 does not solely regulate them. The positive deregulation of Sgcg in qPCR was impressive (about 1,000-fold), but it did not correlate with a higher amount of proteins in dmbo retina. It is likely that Sgcg was strongly regulated at the level of translation, which would explain why this dramatic increase in transcripts had no effect on the protein level, as shown for other genes related to cell adhesion [33]. In addition, the overexpressed γ-sarcoglycan could form aggregates that could be degraded in the endoplasmic reticulum, as supported by the proteolysis process enrichment in MetaCore analysis (Figure 1B). The role of the sgcg gene in the dmbo retinal phenotype remains unclear. Another finding resulting from the microarray analysis was that Hmx1 could act in vivo as a transcriptional repressor or activator. The first in vitro study conducted by Ament et al. showed only a repressor effect, but their work was done in HeLa cells indicating that cellular context may play a role in mediating HMX1 activity [12].

The second part of our study consisted of using our PPM to screen the genome and discover other putative targets of HMX1. More precisely, we focused our attention on identifying HMX1 targets that could be involved in eye development. This would help in understanding the bases of the human oculoauricular syndrome caused by Hmx1 mutation [5]. The first method consisted of using the PPM to screen the RefSeq database of mouse and human. This method retrieved a large but expected number of genes despite the short screening window used. Many of these genes represented interesting candidates (see Appendix 1 for a complete list). To be more selective, we crossed human and mouse selections to keep only genes fitting the PPM in both species. Among the ten genes that we retrieved, four have been reported to be expressed in the eye, during development (EPHA6, MARCKS, and UHRF1), or during adult life (SEPT4) [34-37]. In addition, SH3KBP1 was predicted to be a target of HMX1 and is highly expressed in Schwann cells, where the gene is regulated by SOX10 [38]. Interestingly, a recent study showed that a balance between SOX10 and HMX1 regulates neuronal versus Schwann cell precursor and melanocyte fates [39]. Our second method based on the co-occurrence of HMX1-BS pairs retrieved ten genes in the mouse genome and eight in the human genome. EPHA6, UHRF1, and SH3KBP1 were identified by both methods, increasing the confidence that these three targets were true targets. In the mouse, the Epha6 promoter showed the highest number of HMX1-BS pair combinations with four sites located within [-250,+200]. A wider examination of the region surrounding its TSS showed additional HMX1-BS clusters fitting the PPM in the proximal promoter and in the first intron. With Ptpro and Sema3f, Epha6 belongs to the retinal axon guidance pathway and plays an important role in retinotopic mapping [27,28,34,40,41]. In addition, these three predicted targets occupy key places as inputs of the axon repulsion signaling pathway, supporting a specific and effective action of Hmx1 in this process (see Figure 7 for more details). Finally, the strong and highly significant enrichment of this pathway in the mouse selection obtained with HMX1-BS pairs co-occurrence based method underlined the likely role of Hmx1 in establishing retinal topography. The luciferase assay results provided experimental evidences to validate this hypothesis. For a positive control, we first showed that Hmx1 could act as a repressor and decrease the activity of the Sgcg promoter in the N2A cells. This result was expected based on the dramatic increase of Sgcg expression in dmbo mice. Then, we showed that Hmx1 represses the activity of Ptpro and Sema3f promoters. In the future, we will specifically focus our efforts on the functional study of Epha6, all the more so as the Ephrin pathway was already considered by Schorderet et al. to be a putative target of Hmx1 [5]. We also noted that the human selection contained a new interesting gene expressed in the eye, FOXP1 [42], in addition to UHRF1 and EPHA6. Finally, some of the remaining predicted targets unrelated to the eye could be potential HMX1, NKX2–5, or HMX3 targets in other tissues and will need further investigation.

Several recurrent questions about TFBS identification arose from our study. The first concerns the conservation of CRMs between human and mouse. In fact, we observed poor conservation of our PPM between both species, whereas numerous studies showed that evolutionary conserved regions overlap functional regulatory elements (reviewed in [11]). Actually, approaches integrating comparative genomics data succeed to identify CRMs with a high positive predictive value but overlook CRMs specific for a given species. Single-genome bioinformatic approaches are more convenient to solve this problem. In this manner, a study based on the empirical potential energy of TFs revealed that CRMs occur in conserved and non-conserved regions, and about 55% have a poor conservation score [43]. In particular, this study underlined that the less well-conserved CRMs concern genes related to neural activity. It could be explained by the fact that the nervous system function is specific for species, in contrast with more fundamental processes as transcription, for example. Prediction of neural specific TFBSs appears to be harder than others; fortunately, it has been shown that homotypic CRMs are a good predictor of regulatory elements, especially for target genes related to TFs involved in neural development [17]. In a general way, homotypic CRMs are strongly associated with the region surrounding the TSS and the developmental enhancers, with no systematic phylogenetic conservation. These observations rationalize the results we obtained with a single-genome approach and the PPM-based method focused on the co-occurrence of HMX1-BS pairs, in particular with the mouse axon guidance pathway enrichment. Finally, we think that adding a distance range constraint to the PPM gave more accurate results than simply counting the TFBSs in the screening window. Actually, optimal distances for interactions are supposed to be specific for a given TF and including this parameter in the model should give more specific results [26].

In conclusion, our strategy was successful in identifying HMX1 targets because the PPM we constructed based on the P15 microarray data revealed, a posteriori, an important pathway involved in retinal development (I.E., axon repulsion during retinal axon guidance) in addition to the first three targets (I.E., Sgcg, Tshz2, and Slc6a9). These subsequent outcomes provided additional proof of the robustness of our PPM approach, and open up new opportunities to focus experimental investigations on this specific aspect of the Hmx1 pathway.

Appendix 1. Official gene symbols and gene accession numbers of all predicted targets.

Appendix 2. Sequences and amplicon sizes of qPCR primer pairs used in this study.

Appendix 3 Upregulated genes with a FDR<0.1 and a fold-change>1.2.

Appendix 4. Downregulated genes with a FDR<0.1 and a fold-change>1.2.

Appendix 5. Gene functional classification of FDR<0.1 differentially expressed genes according to DAVID software.

Acknowledgments

We thank the Center for Integrative Genomics of the University of Lausanne for performing microarray analysis. This work was supported by grants n°31003A-124990 and 31003A-143474 from the Swiss National Science Foundation. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  1. Stadler HS, Murray JC, Leysens NJ, Goodfellow PJ, Solursh M. Phylogenetic conservation and physical mapping of members of the H6 homeobox gene family. Mamm Genome. 1995; 6:383-8. [PMID: 7647458]
  2. Wang W, Lo P, Frasch M, Lufkin T. Hmx: an evolutionary conserved homeobox gene family expressed in the developing nervous system in mice and Drosophila. Mech Dev. 2000; 99:123-37. [PMID: 11091080]
  3. Wang W, Lufkin T. Hmx homeobox gene function in inner ear and nervous system cell-type specification and development. Exp Cell Res. 2005; 306:373-9. [PMID: 15925593]
  4. Yoshiura K, Leysens NJ, Reiter RS, Murray JC. Cloning, characterization, and mapping of the mouse homeobox gene Hmx1. Genomics. 1998; 50:61-8. [PMID: 9628823]
  5. Schorderet DF, Nichini O, Boisset G, Polok B, Tiab L, Mayeur H, Raji B, De la Houssaye G, Abitbol MM, Munier FL. Mutation in the human homeobox gene NKX5–3 causes an oculo-auricular syndrome. Am J Hum Genet. 2008; 82:1178-84. [PMID: 18423520]
  6. Munroe RJ, Prabhu V, Acland GM, Johnson KR, Harris BS, O’Brien TP, Welsh IC, Noden DM, Schimenti JC. Mouse H6 Homeobox 1 (Hmx1) mutations cause cranial abnormalities and reduced body mass. BMC Dev Biol. 2009; 9:27 [PMID: 19379485]
  7. Quina LA, Tempest L, Hsu Y-WA, Cox TC, Turner EE. Hmx1 is required for the normal development of somatosensory neurons in the geniculate ganglion. Dev Biol. 2012; 365:152-63. [PMID: 22586713]
  8. Quina LA, Kuramoto T, Luquetti DV, Cox TC, Serikawa T, Turner EE. Deletion of a conserved regulatory element required for Hmx1 expression in craniofacial mesenchyme in the dumbo rat: a newly identified cause of congenital ear malformation. Dis Model Mech. 2012; 5:812-22. [PMID: 22736458]
  9. Shelest E, Wingender E. Construction of predictive promoter models on the example of antibacterial response of human epithelial cells. Theor Biol Med Model. 2005; 2:2 [PMID: 15647113]
  10. Werner T, Fessele S, Maier H, Nelson PJ. Computer modeling of promoter organization as a tool to study transcriptional coregulation. FASEB J. 2003; 17:1228-37. [PMID: 12832287]
  11. Hardison RC, Taylor J. Genomic approaches towards finding cis-regulatory modules in animals. Nat Rev Genet. 2012; 13:469-83. [PMID: 22705667]
  12. Amendt BA, Sutherland LB, Russo AF. Transcriptional antagonism between Hmx1 and Nkx2.5 for a shared DNA-binding site. J Biol Chem. 1999; 274:11635-42. [PMID: 10206974]
  13. Warren SA, Terada R, Briggs LE, Cole-Jeffrey CT, Chien W-M, Seki T, Weinberg EO, Yang TP, Chin MT, Bungert J, Kasahara H. Differential role of Nkx2–5 in activation of the atrial natriuretic factor gene in the developing versus failing heart. Mol Cell Biol. 2011; 31:4633-45. [PMID: 21930795]
  14. Kasahara H, Usheva A, Ueyama T, Aoki H, Horikoshi N, Izumo S. Characterization of homo- and heterodimerization of cardiac Csx/Nkx2.5 homeoprotein. J Biol Chem. 2001; 276:4570-80. [PMID: 11042197]
  15. Lee Y, Shioi T, Kasahara H, Jobe SM, Wiese RJ, Markham BE, Izumo S. The cardiac tissue-restricted homeobox protein Csx/Nkx2.5 physically associates with the zinc finger protein GATA4 and cooperatively activates atrial natriuretic factor gene expression. Mol Cell Biol. 1998; 18:3120-9. [PMID: 9584153]
  16. Ryu J-R, Najand N, Brook WJ. Tinman is a direct activator of midline in the Drosophila dorsal vessel. Dev Dyn. 2011; 240:86-95. [PMID: 21108319]
  17. Gotea V, Visel A, Westlund JM, Nobrega MA, Pennacchio LA, Ovcharenko I. Homotypic clusters of transcription factor binding sites are a key component of human promoters and enhancers. Genome Res. 2010; 20:565-77. [PMID: 20363979]
  18. Smyth GK. Linear models and empirical bayes methods for assessing differential expression in microarray experiments. Stat Appl Genet Mol Biol. 2004; 3:Article3 [PMID: 16646809]
  19. Benjamini Y, Hochberg Y. Controlling the false discovery rate: a practical and powerful approach to multiple testing. J R Stat Soc Series B Stat Methodol. 1995; 57:289-300.
  20. Huang DW, Sherman BT, Lempicki RA. Systematic and integrative analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc. 2009; 4:44-57. [PMID: 19131956]
  21. Huang DW, Sherman BT, Lempicki RA. Bioinformatics enrichment tools: paths toward the comprehensive functional analysis of large gene lists. Nucleic Acids Res. 2009; 37:1-13. [PMID: 19033363]
  22. Siegert S, Cabuy E, Scherf BG, Kohler H, Panda S, Le Y-Z, Fehling HJ, Gaidatzis D, Stadler MB, Roska B. Transcriptional code and disease map for adult retinal cell types. Nat Neurosci. 2012; 15:487-95. [PMID: 22267162]
  23. Vinson C, Chatterjee R, Fitzgerald P. Transcription factor binding sites and other features in human and Drosophila proximal promoters. Subcell Biochem. 2011; 52:205-22. [PMID: 21557085]
  24. Bernard V, Lecharny A, Brunaud V. Improved detection of motifs with preferential location in promoters. Genome. 2010; 53:739-52. [PMID: 20924423]
  25. Fort P, Estrada F-J, Bordais A, Mornet D, Sahel J-A, Picaud S, Vargas HR, Coral-Vázquez RM, Rendon A. The sarcoglycan-sarcospan complex localization in mouse retina is independent from dystrophins. Neurosci Res. 2005; 53:25-33. [PMID: 15993965]
  26. Hu Z, Hu B, Collins JF. Prediction of synergistic transcription factors by function conservation. Genome Biol. 2007; 8:R257 [PMID: 18053230]
  27. Stepanek L, Sun QL, Wang J, Wang C, Bixby JL. CRYP-2/cPTPRO is a neurite inhibitory repulsive guidance cue for retinal neurons in vitro. J Cell Biol. 2001; 154:867-78. [PMID: 11514594]
  28. Shintani T, Ihara M, Sakuta H, Takahashi H, Watakabe I, Noda M. Eph receptors are negatively controlled by protein tyrosine phosphatase receptor type O. Nat Neurosci. 2006; 9:761-9. [PMID: 16680165]
  29. Tian N. Visual experience and maturation of retinal synaptic pathways. Vision Res. 2004; 44:3307-16. [PMID: 15535998]
  30. Freeman NE, Templeton JP, Orr WE, Lu L, Williams RW, Geisert EE. Genetic networks in the mouse retina: growth associated protein 43 and phosphatase tensin homolog network. Mol Vis. 2011; 17:1355-72. [PMID: 21655357]
  31. Santos JS, Fonseca NA, Vieira CP, Vieira J, Casares F. Phylogeny of the teashirt-related zinc finger (tshz) gene family and analysis of the developmental expression of tshz2 and tshz3b in the zebrafish. Dev Dyn. 2010; 239:1010-8. [PMID: 20108322]
  32. Reed BT, Sullivan SJ, Tsai G, Coyle JT, Esguerra M, Miller RF. The glycine transporter GlyT1 controls N-methyl-D-aspartic acid receptor coagonist occupancy in the mouse retina. Eur J Neurosci. 2009; 30:2308-17. [PMID: 20092573]
  33. Schwanhäusser B, Busse D, Li N, Dittmar G, Schuchhardt J, Wolf J, Chen W, Selbach M. Global quantification of mammalian gene expression control. Nature. 2011; 473:337-42. [PMID: 21593866]
  34. Carreres MI, Escalante A, Murillo B, Chauvin G, Gaspar P, Vegar C, Herrera E. Transcription factor Foxd1 is required for the specification of the temporal retina in mammals. J Neurosci. 2011; 31:5673-81. [PMID: 21490208]
  35. Zolessi FR, Arruti C. MARCKS in advanced stages of neural retina histogenesis. Dev Neurosci. 2004; 26:371-9. [PMID: 15855766]
  36. Tittle RK, Sze R, Ng A, Nuckels RJ, Swartz ME, Anderson RM, Bosch J, Stainier DYR, Eberhart JK, Gross JM. Uhrf1 and Dnmt1 are required for development and maintenance of the zebrafish lens. Dev Biol. 2011; 350:50-63. [PMID: 21126517]
  37. Pache M, Zieger B, Bläser S, Meyer P. Immunoreactivity of the septins SEPT4, SEPT5, and SEPT8 in the human eye. J Histochem Cytochem. 2005; 53:1139-47. [PMID: 15923366]
  38. Hodonsky CJ, Kleinbrink EL, Charney KN, Prasad M, Bessling SL, Jones EA, Srinivasan R, Svaren J, McCallion AS, Antonellis A. SOX10 regulates expression of the SH3-domain kinase binding protein 1 (Sh3kbp1) locus in Schwann cells via an alternative promoter. Mol Cell Neurosci. 2012; 49:85-96. [PMID: 22037207]
  39. Adameyko I, Lallemend F, Aquino JB, Pereira JA, Topilko P, Müller T, Fritz N, Beljajeva A, Mochii M, Liste I, Usoskin D, Suter U, Birchmeier C, Ernfors P. Schwann cell precursors from nerve innervation are a cellular origin of melanocytes in skin. Cell. 2009; 139:366-79. [PMID: 19837037]
  40. Bevins N, Lemke G, Reber M. Genetic dissection of EphA receptor signaling dynamics during retinotopic mapping. J Neurosci. 2011; 31:10302-10. [PMID: 21753007]
  41. Claudepierre T, Koncina E, Pfrieger FW, Bagnard D, Aunis D, Reber M. Implication of neuropilin 2/semaphorin 3F in retinocollicular map formation. Dev Dyn. 2008; 237:3394-403. [PMID: 18942144]
  42. Cheng L, Chong M, Fan W, Guo X, Zhang W, Yang X, Liu F, Gui Y, Lu D. Molecular cloning, characterization, and developmental expression of foxp1 in zebrafish. Dev Genes Evol. 2007; 217:699-707. [PMID: 17876603]
  43. Yu X, Lin J, Zack DJ, Qian J. Identification of tissue-specific cis-regulatory modules based on interactions between transcription factors. BMC Bioinformatics. 2007; 8:437 [PMID: 17996093]