Chemokine Signaling Mediates Self-Organizing Tissue Migration in the Zebrafish Lateral Line

Similar documents
Michal Reichman-Fried

The octavolateralis sensory system of an aquatic vertebrate

7.22 Final Exam points

Supplement Figure S1:

Neural Induction. Chapter One

T H E J O U R N A L O F C E L L B I O L O G Y

7.22 Example Problems for Exam 1 The exam will be of this format. It will consist of 2-3 sets scenarios.

Supplementary Figure 1

To assess the localization of Citrine fusion proteins, we performed antibody staining to

Lecture 20: Drosophila embryogenesis

Supplemental Information. Boundary Formation through a Direct. Threshold-Based Readout. of Mobile Small RNA Gradients

Neural Development. How does a single cell make a brain??? How are different brain regions specified??? Neural Development

Aquatic Vertebrates Platform Services (See services description below, page 3)

Supplemental Fig. 1. Mcr alleles show defects in tracheal tube size and luminal protein accumulation. (A-F) Confocal projections of living stage 15

Mesoderm Formation. Fate map of early gastrula. Only two types of mesoderm are induced. Mesoderm induction by the vegetal hemisphere

Supplementary Figure 1. Homozygous rag2 E450fs mutants are healthy and viable similar to wild-type and heterozygous siblings.

Lef1 regulates Dusp6 to influence neuromast formation and spacing in the zebrafish posterior lateral line primordium

Supplemental Information

Lecture 3 MOLECULAR REGULATION OF DEVELOPMENT

CS262 Lecture 12 Notes Single Cell Sequencing Jan. 11, 2016

Lecture 20: Drosophila melanogaster

Xenopus gastrulation. Dorsal-Ventral Patterning - The Spemann Organizer. Two mesoderm inducing signals

-RT RT RT -RT XBMPRII

Reading. Lecture III. Nervous System Embryology. Biology. Brain Diseases. September 5, Bio 3411 Lecture III. Nervous System Embryology

Lecture III. Nervous System Embryology

Paquet et al Supplementary Information

1. Supplementary Tables

Supporting Information

Readings. Lecture IV. Mechanisms of Neural. Neural Development. September 10, Bio 3411 Lecture IV. Mechanisms of Neural Development

HOW DOES Wnt SIGNALING POSTERIORIZE THE NEURAL PLATE?

Pregnenolone Stabilizes Microtubules and Promotes Zebrafish Embryonic Cell Movement

7.013 Practice Quiz

Genome manipulation by homologous recombination in Drosophila Xiaolin Bi and Yikang S. Rong Date received (in revised form): 9th May 2003

Exam 1 ID#: October 1, 2006

Nature Biotechnology: doi: /nbt Supplementary Figure 1

Trasposable elements: Uses of P elements Problem set B at the end

Non-coding Function & Variation, MPRAs II. Mike White Bio /5/18

Lecture 8: Transgenic Model Systems and RNAi

Additional Practice Problems for Reading Period

Activity 47.1 What common events occur in the early development of animals? 1. What key events occur at each stage of development?

GENETICS - CLUTCH CH.15 GENOMES AND GENOMICS.

Supplementary Methods

fga phenotypes were also analyzed in the tracheal system, the organ responsible for

A CRISPR/Cas9 Vector System for Tissue-Specific Gene Disruption in Zebrafish

Supplementary Information. Isl2b regulates anterior second heart field development in zebrafish

Later Development. Caenorhabditis elegans. Later Processes 10/06/12. Cytoplasmic Determinants Fate Mapping & Cell Fate Limb Development

Developmental Biology. Cell Fate, Potency, and Determination

Cell Lineage in the Development of the Leech Nervous System

ksierzputowska.com Research Title: Using novel TALEN technology to engineer precise mutations in the genome of D. melanogaster

Supporting Online Material for

SUPPLEMENTARY INFORMATION

Regulation of Acetylcholine Receptor Clustering by ADF/Cofilin- Directed Vesicular Trafficking

SUPPLEMENTARY INFORMATION

a) Stock 4: 252 flies total: 120 are P[w+]: 64 are CyO; 56 are TM3; all are females. 132 are w-: 68 are CyO; 64 are TM3; all are males.

What s the most complex problem in biology?

Chapter 47. Development

(a) Scheme depicting the strategy used to generate the ko and conditional alleles. (b) RT-PCR for

Optimizing Cas9 and sgrna concentrations for increasing mutation efficiency. Nature Methods: doi: /nmeth.3360

Regulation of Hematopoietic Stem Cells and their Bone Marrow Niches by the Coagulation System*.

A Survey of Genetic Methods

Supplementary Figure 1. Espn-1 knockout characterization. (a) The predicted recombinant Espn-1 -/- allele was detected by PCR of the left (5 )

Sydney Brenner: Present

Mitotic cell rounding and epithelial thinning regulate lumen

Quantitative real-time RT-PCR analysis of the expression levels of E-cadherin

LIFE Formation III. Morphogenesis: building 3D structures START FUTURE. How-to 2 PROBLEMS FORMATION SYSTEMS SYSTEMS.

SUPPLEMENTARY INFORMATION

Nature Biotechnology: doi: /nbt.4166

SUPPLEMENTARY INFORMATION

Understanding embryonic head development. ANAT2341 Tennille Sibbritt Embryology Unit Children s Medical Research Institute

Hemidesmosome. Focal adhesion

Supplementary Figures

JCB. Supplemental material THE JOURNAL OF CELL BIOLOGY. Paul et al.,

7.012 Problem Set 5. Question 1

sides of the aleurone (Al) but it is excluded from the basal endosperm transfer layer

Chapter 6 - Molecular Genetic Techniques

SUPPLEMENTARY INFORMATION

Biology 4361 Developmental Biology Lecture 4. The Genetic Core of Development

ANAT2341 Embryology Introduction: Steve Palmer

Characterization of a novel gene involved in border cell migration

Gene Expression: Transcription

Cell position and developmental fate in leech embryogenesis

(B) Comparable expression of major integrin subunits and glycoproteins on the surface of resting WT and Lnk -/- platelets.

John Gurdon was testing the hypothesis of genomic equivalence or that when cells divide they retain a full genomic compliment.

APPLICATION SPECIFIC PROTOCOL CELL MIGRATION FOR ADHERENT CELLS

Sdf1/Cxcr4 signaling controls the dorsal migration of endodermal cells during zebrafish gastrulation

Schematic representation of the endogenous PALB2 locus and gene-disruption constructs

Drosophila White Paper 2003 August 13, 2003

(A) Schematic depicting morphology of cholinergic DA and DB motor neurons in an L1

Fig. S1. Molecular phylogenetic analysis of AtHD-ZIP IV family. A phylogenetic tree was constructed using Bayesian analysis with Markov Chain Monte

GM130 Is Required for Compartmental Organization of Dendritic Golgi Outposts

Supporting Information

Sperm cells are passive cargo of the pollen tube in plant fertilization

Supplementary Figure 1. Phenotype, morphology and distribution of embryonic GFP +

Enzyme that uses RNA as a template to synthesize a complementary DNA

7 Gene Isolation and Analysis of Multiple

Supplemental Information. Glutamylation Regulates Transport, Specializes. Function, and Sculpts the Structure of Cilia

Modeling Periodic Aspect of Limb Pattern Formation

SCREENING AND PRESERVATION OF DNA LIBRARIES

λ N -GFP: an RNA reporter system for live-cell imaging

Guided differentiation of ES cell into mesenchymal stem cell

Transcription:

Developmental Cell 10, 673 680, May, 2006 ª2006 Elsevier Inc. DOI 10.1016/j.devcel.2006.02.019 Chemokine Signaling Mediates Self-Organizing Tissue Migration in the Zebrafish Lateral Line Short Article Petra Haas 1 and Darren Gilmour 1, * 1 European Molecular Biology Laboratory Meyerhofstrasse 1 Heidelberg 69117 Germany Summary The shape of most complex organ systems arises from the directed migration of cohesive groups of cells. Here, we dissect the role of the chemokine guidance receptor Cxcr4b in regulating the collective migration of one such cohesive tissue, the zebrafish lateral line primordium. Using in vivo imaging, we show that the shape and organization of the primordium is surprisingly labile, and that internal cell movements are uncoordinated in embryos with reduced Cxcr4b signaling. Genetic mosaic experiments reveal that single cxcr4b mutant cells can migrate in a directional manner when placed in wild-type primordia, but that they are specifically excluded from the leading edge. Moreover, a remarkably small number of SDF1a-responsive cells are able to organize an entire cxcr4b mutant primordium to restore migration and organogenesis in the lateral line. These results reveal a role for chemokine signaling in mediating the self-organizing migration of tissues during morphogenesis. Introduction The directed migration of cells or cellular processes sculpts the three-dimensional form of most complex organs. Genetic screens have led to the identification of a large number of secreted proteins and their cognate receptors, which are required for the guidance of migrating cells during morphogenesis, including proteins of the FGF, EGF, and chemokine families (Affolter and Weijer, 2005; Mine et al., 2005; Tran and Miller, 2003). While significant progress has been made in identifying the factors determining the direction in which cells migrate, the mechanisms that ensure coordinated cell movements remain elusive. Such coordination is essential during organogenesis, during which every cell must adopt a position that is appropriate for its function. One clue to how these precise movements may be choreographed comes from the observation that cells undergoing organogenesis are rarely found as chemotactic individuals, but rather migrate collectively in the form of cohesive groups or tissues, such as sheets, chains, or clusters. Such collective migration behavior is observed during the formation of the vasculature, respiratory, and nervous systems and is adopted by many tumors during metastasis (Friedl, 2004). Given the prevalence of such migrating tissues in vivo, it is surprising that little is known about the cellular basis of their organization. *Correspondence: gilmour@embl.de For example, in most cases the extent to which extrinsic cues shape these moving groups or control cell movements within them is completely unclear. The development of the zebrafish lateral line system offers several features, such as simple in vivo imaging and genetic tractability, which make it an attractive model system for studying collective cell migration behavior during organogenesis (Ghysen and Dambly-Chaudiere, 2004). It is comprised of a series of mechanosensory hair cell organs (neuromasts) that are deposited throughout the skin by the posterior lateral line primordium (pllp), a cohesive mass of more than 100 migrating cells. The path followed by the primordium is defined by the expression of a zebrafish homolog of the chemokine stromal-derived factor 1 (SDF1), which the primordium detects through the expression of its receptor Cxcr4b, and knockdown of either the ligand or receptor results in a similar strong defect in pllp migration (David et al., 2002; Li et al., 2004). This chemokine-receptor pair, which was first identified for its role in leukocyte homing (Peled et al., 1999), is known to direct the migration of a number of cell types, including neurons (Tran and Miller, 2003), primordial germ cells (Doitsidou et al., 2002; Knaut et al., 2003), and neural crest cells (Belmadani et al., 2005), during development. Here, we combine genetic mosaic analysis with in vivo imaging to address the role of Cxcr4/SDF1 signaling in organizing the lateral line primordium. Our results shed light on the dynamics of concerted cell movements during organogenesis. Results The Lateral Line Primordium Is an Organized Cohort of Migrating Cells In order to image morphogenesis of the lateral line, we generated transgenic zebrafish lines in which a membrane-tethered version of GFP (Koster and Fraser, 2001) is placed under the control of the Claudin B promoter. Claudin B is a member of the tetraspanin family of tight junction proteins that is transcribed at high levels in organs derived from sensory placodes, including the lateral line (Kollmar et al., 2001; Lopez-Schier et al., 2004). As shown in Figure 1, 8 kb of the Claudin B promoter are sufficient to drive the expression of GFP in every cell of the migrating pllp, as well as the neuromast organs, the progenitors of which can be identified as 2 3 rosette-like structures in the migrating primordium, and a chain of interneuromast cells that it lays down en route (Figure 1C; Movie S1 in the Supplemental Data available with this article online). We first used this transgenic line (cldnbgfp) to address how many cells of the pllp display filopodia and pseudopodia characteristic of active migration. Such cellular extensions are often taken as an indicator of which cells within moving tissues are guiding directed migration (Fulga and Rorth, 2002; Ribeiro et al., 2002). As expected, dynamic filopodia and pseudopodia are most apparent in the 2 4 cells at the very tip of the pllp (Figure 1D; Movie S2). Surprisingly, however, these processes are also exhibited by the

Developmental Cell 674 Figure 1. Capturing Cell Behavior within the pllp by Using CldnBGFP (A and B) CldnBGFP transgenic embryos recapitulate the endogenous Claudin B expression pattern and show GFP expression in the migrating primordium, deposited neuromasts, and connecting interneuromast cells. Other expression domains include pronephros (arrows) and skin. The scale bar is 20 mm. (C) An overview of a time-lapse movie showing 10 hr of lateral line morphogenesis with CldnBGFP. The lateral line primordium migrates at a speed of w66 mm/hr at 25ºC. Forming proneuromasts at the trailing edge decelerate, causing the tissue to stretch, before being deposited (Movie S1). The scale bar is 100 mm. (D and E) Many cells of the primordium display cellular extensions (compare [D] to [E]; Movies S2 and S3). (F) Kymograph analysis of cell movement. In this kymograph from wild-type, the traces are predominantly parallel, meaning that cells within the primordium migrate at a constant speed and maintain a relative position (dotted lines on the right). The deceleration of cells at the back of the primordium can be seen in the increased distance between the two dotted lines on the left (Movie S4). (G) Cells within cxcr4b mutant primordium continue to move back and fourth, as revealed by a zig-zag pattern of traces in the kymograph (Movie S5). The scale bar is 20 mm. majority of cells running along the sides of the migrating tissue, giving it a millipede-like appearance (Figure 1E; Movie S3). Based on this morphological criterion alone, it appears that many cells of the primordium respond directly to extrinsic guidance cues. Cxcr4 Signaling Is Required for Coordinated Motility within the pllp In order to address the role of Cxcr4b signaling in organizing cell behaviors within this migrating group, we crossed the cldnbgfp transgene into the cxcr4b mutant background. The normal tissue polarity is strongly affected in the absence of Cxcr4b, and the pllp is significantly rounder than in wild-type embryos (Figures 1G and 4E; Movie S5). Time-lapse analysis shows that internal cell movements are uncoordinated in cxcr4b mutants (compare kymographs in Figures 1F and 1G). Cells of the pllp appear to have the potential to migrate in random directions in the absence of Cxcr4b function. However, as they are held together in this context, such uncoordinated movements cancel each other out and result in greatly reduced net displacement of the cell mass. In order to allow migration of a tissue such as the pllp, it is essential that motile cells are orientated in the same direction, which requires Cxcr4b-mediated detection of a stripe of SDF1a. SDF1a Allows Bidirectional Migration along the Lateral Line Chemokines such as SDF1 are the archetypal vertebrate chemoattractants (Rossi and Zlotnik, 2000). Not only has SDF1 been shown to guide cells in a concentration-dependent manner in a number of different species, but the misexpression of SDF1 in zebrafish embryos can affect the migration of both PGCs and the lateral line primordium (Doitsidou et al., 2002; Li et al., 2004). These ectopic expression experiments show that SDF1 can act as a chemoattractant in vivo and suggest that the

Cxcr4-Mediated Tissue Migration 675 Figure 2. Bidirectional Migration of the pllp along an SDF1a Stripe (A) Expression of SDF1a in the trunk of wild-type (upper panel) and fss mutant siblings (lower panel). In fss, the stripe of SDF1a followed by the pllp primordium is truncated, whereas a ventral expression domain at the level of the pronephros remains unaffected. (B) Double labeling for SDF1a mrna (green) and the pll nerve (aactub, red) in wild-type and fss embryos. The lateral line makes a sharp turn at the point at which SDF1a expression fades in fss (lower panel). (C) Diagram indicating major routes taken by pllp in fss mutants. (D) Example of a primordium being attracted by a ventral source of SDF1a in fss. (E) Time-lapse movie showing the pllp undergoing a U-turn maneuver. The upper start panel shows a rounded primordium; a small group of cells projects backward, causing the tissue to rotate. Once this U-turn is complete, the pllp readopts its normal polarized morphology and migrates at normal speed in the reverse direction and even deposits a proneuromast (Movie S6). lateral line primordium normally moves up a gradient of this chemokine. On the other hand, previous grafting experiments carried out in Axolotl have shown that 180º rotation of the epidermis directly in front of the primordium appeared to have little effect on the directionality of its migration (Smith et al., 1990). While these and other embryological studies (Stone, 1923) argue against the presence of a graded signal in this species, they do not rule out the possibility that a gradient was reestablished after surgery. In order to determine which of these two potential modes of chemokine-mediated guidance are employed during primordium migration, we addressed the behavior of the lateral line in mutants with changes in the endogenous SDF1a expression domain. Embryos mutant for fused somites/tbx 24 (fss) (Nikaido et al., 2002) show a very interesting alteration in the expression of the lateral line stripe of SDF1a, which is present in the anterior 5 10 somites but is absent further posteriorly, presumably due to somite patterning defects in this mutant (Figure 2A). In a significant proportion of fss mutants, the pllp makes a U-turn maneuver and migrates backward along the same stripe of SDF1a it previously followed (11.8%, 27/229; Figures 2B and 2C). Here, the primordium migrates normally and even deposits neuromasts; however, it now moves in the reverse direction (Figure 2E; Movie S6). This demonstration that the primordium treats this path of SDF1a like a two-way street suggests that the normal head-to-tail directionality of lateral line migration is not determined by a polarized distribution of the chemokine guidance cue, but rather by the organization of the primordium itself. Interestingly, embryos injected with morpholinos against N-cadherin, which is highly expressed by the primordium, show similar pathfinding phenotypes; however, the expression of SDF1a in these embryos has not been addressed (Kerstetter et al., 2004). Collective Migration within the Lateral Line We next addressed the extent to which Cxcr4b/SDF1a signaling controls individual cell behaviors within this migrating tissue by carrying out a series of genetic mosaic experiments via cell transplantation. We first tested this approach by transplanting rhodamine dextranlabeled wild-type cells into wild-type cldnbgfp transgenic embryos at the blastula stage. The recipient embryos were allowed to develop to 36 hpf before screening for the presence of transplanted cells in the pllp. While the positioning of cells cannot be precisely controlled, we could efficiently generate embryos with mosaic pllps; greater than 10% transplanted host embryos showed the presence of 5 10 red donor cells on average (16/136). Their position was mapped by plotting a red fluorescence intensity profile of the leading 150 mm of the tissue (Figure 3B). Importantly, when many such plots are superimposed, it becomes clear that transplanted wild-type cells are scattered throughout the host primordium and show no clear bias in the positions they occupy. We next addressed the role of Cxcr4b in controlling individual cell behaviors by transplanting homozygous mutant cells into wild-type cldnbgfp transgenic fish. As shown in Figure 3D, cells lacking the chemokine guidance receptor move in a directed manner when placed in a wild-type primordium; thus, they can be guided by Cxcr4b-independent interactions with neighboring cells. Time-lapse and kymograph analysis shows that these mutant cells migrate at the same

Developmental Cell 676 Figure 3. Cell-Cell Interactions Regulate Migration in the Lateral Line Primordium (A) Rhodamine-labeled wild-type cells transplanted into wild-type embryos transgenic for cldnbgfp (upper panel). The lower panels show an example of a mosaic primordium in which donor cells colonize the tip (middle; rhodamine only, lower; rhodamine plus cldnbgfp). (B) Plot of red intensity profile from 16 such transplants, showing maximum (blue) and minimum (red). The green trace shows a profile of primordium from (A). (C) CldnbGFP wild-type cells transplanted into an mrfp-labeled wild-type primordium. The right panel shows a magnified view of internal cell clones displaying filopodia in the direction of migration (arrowhead). No such projections are observed from the rear of the cell (arrows). Asterisks mark the lateral line nerve from the donor, which also expresses cldnbgfp (Movie S7). (D and E) Mutant cxcr4b cells migrate normally when transplanted into wild-type primordia, but they adopt trailing positions. Intensity profiling shows that in no case (n = 21) were cxcr4b cells present in the leading 20 mm of the primordium. (F) The right panel shows a magnified cxcr4b mutant clone transplanted into an mrfplabeled wild-type primordium. Only the leading edges display filopodial protrusions (arrowhead, Movie S8). (G) Two-color time-lapse of a mosaic primordium, in which red indicates cxcr4b cells and the dotted line represents the region used for the kymograph (Movie S9). (H) Kymograph time-lapse movie from (G). Red and green traces are parallel, indicating that cxcr4b cells migrate at the same speed as their wild-type counterparts. speed and maintain a constant position in relation to their neighbors, suggesting that the precise positioning of cells within the pllp is not based on Cxcr4b activity (Figures 3G and 3H; Movie S9). However, mutant cells show one clear difference when compared to wildtype donor cells: they never occupy positions at the leading edge in wild-type hosts. We were able to restrict the region that remains free of mutant cells to the first 20 microns at the very tip of the migrating primordium, an area that corresponds approximately to one leading cell diameter (Figure 3E). Consistent with this inability to assume the leading position, in no case was the presence of mutant clones, no matter how large, able to disrupt the migration of the wild-type primordium (n > 50). As mutant cells presumably occupy random positions initially within the group, this suggests that they are actively excluded from the leading edge. Are cells within the primordium actively migrating or simply passively carried by motile neighbors? In order to answer this question, we imaged the morphology of cells within the primordium by transplanting cells from cldnbgfp transgenics into embryos in which all cells were labeled red by the injection of membrane RFP (mrfp) mrna (Gong et al., 2004). This approach shows that single wild-type cells further back within a wild-type primordium appear polarized and project highly dynamic extensions in the direction of migration (Figure 3C; Movie S7). Interestingly, Cxcr4b-deficient cells have

Cxcr4-Mediated Tissue Migration 677 Figure 4. Cxcr4-Positive Cells Act as a Lateral Line Organizing Center (A) CldnbGFP wild-type cells do not actively translocate to the leading edge of an mrfp-labeled cxcr4b primordium. Wild-type cells (two middle panels) point randomly (arrows) and display no lateral displacement over a period of 132 min. Kymograph analysis confirms that wild-type cells tumble with mutant neighbors. (B) Wild-type cells at the tip of the primordium project in the direction of migration, pulling mutant cells with them. Initially, this polarizing influence has a limited range, causing the tissue to stretch (the dotted line highlights the separation of red-labeled wild-type cells) (Movies S10 S12). (C) Kymograph analysis shows that it takes time for migration in these mutant primordia to become coordinated. The top half of this kymograph resembles that shown for cxcr4b mutants, with zig-zag lines, whereas, in the bottom half, the lines are parallel as in wild-type (Movie S13). (D) A red intensity profile from mosaic primordia shows that wild-type cells always colonize the tip of the migrating primordium (lower panel, n = 17). The peak in minimum intensity reading between 130 and 150 mm demonstrates the narrow region that is occupied by wild-type cells in every sample. (E) The shape of mutant primordia is also rescued by the presence of wild-type cells (roundness = 4pA/P 2, where A = area and P = perimeter). Error bars show standard deviation. (F) The plot shows the distance traveled by wild-type (blue), rescued (red), and cxcr4b mutant (green) primordia by 40 hpf. Rescued primordia trail wild-type by 60 500 mm, a time delay of 1 7.5 hr. (G) Comparison of a wild-type, cxcr4b, and rescued embryo at 42 hpf. The deposition of lateral line neuromasts in the rescued sample is indistinguishable from that in a wild-type embryo. a similar dynamic behavior when transplanted into a wild-type primordium, suggesting that it is not a direct response to the extrinsic SDF1a cue, but is rather induced by interactions with neighboring cells (Figure 3F; Movie S8). A Cxcr4b-Dependent Organizing Center of the Lateral Line We next asked to what extent, if any, transplantation of wild-type cells could rescue the phenotype of Cxcr4bdeficient pllps. We found that transplantation of even small numbers of wild-type cells is able to rescue the directed migration of otherwise mutant primordia with remarkable efficiency. As few as 4 wild-type cells can rescue the migration of more than 100 mutant neighbors, demonstrating that fewer than 5% of the cells must express the chemokine guidance receptor to allow directional migration. Furthermore, as the primordium guides the extending lateral line nerve (Gilmour et al., 2004), which in turn guides associated migrating glial precursors (Gilmour et al., 2002), the presence of a small group of Cxcr4b-expressing cells is sufficient to rescue the migration of all three cell types (data not shown). We find that only the first 20 mm of the primordium harbor wild-type cells in all rescued samples, the same sized region that remains free of donor cells in mutant into wildtype transplants (Figure 4D). Therefore, through a series of complementary genetic mosaic experiments, we have unequivocally identified the minimum region of the primordium that must express Cxcr4b in order to ensure the guided migration of the lateral line primordium. What is very surprising from our findings is the efficiency with which transplantation of a small number of wild-type cells can rescue the lateral line migration defect in cxcr4b mutant embryos. In almost nine out of ten samples in which wild-type cells are present in mutant primordia, they are found at the tip and rescue migration (85%, 94/110). This rescue rate is significantly higher than that expected from the random distribution of transplanted cells; in our original control transplants, the leading 20 mm of wild-type host primordia showed the presence of wild-type donor cells in fewer than

Developmental Cell 678 40% of the embryos (6/16). The efficiency with which wild-type cells colonize the tip of mutants is all the more remarkable when we consider that a significant proportion of cells of the primordium are becoming incorporated into proneuromasts at this stage. Indeed, in the small fraction of embryos in which wild-type cells do not rescue, they are invariably incorporated into forming neuromast organs, which may explain their lack of mobility (see rosette-like structures in Figure 1C; Movie S1). How do these wild-type cells colonize the tip of cxcr4b mutant pllps so efficiently? The most obvious explanation is that these cells detect the SDF1a cue and actively translocate through the mutant tissue toward the leading edge. To address this possibility, we examined the behavior of wild-type cells in mutant primordia by transplantation of wild-type cldngfp cells into mrfp-labeled cxcr4b mutants. This revealed that while internal wildtype cells appeared motile and polarized, they do not persistently migrate in the direction of the SDF1a source and do not oppose the tumbling movement of the host primordia (Figure 4A; Movies S10 and S11). This apparent inability to actively translocate is consistent with the observation that Cxc4b-expressing cells are often distributed along the entire length of rescued primordia. Kymograph analysis shows that they do not change position with respect to their mutant neighbors, arguing against the idea that the internal positioning of cells within the group is determined by Cxcr4b expression (Figure S1). An alternative interpretation for the high efficiency of rescue is based on the morphological plasticity of this migrating tissue. Imaging of mosaic pllps during earlier stages before rescue shows that they are essentially indistinguishable from Cxcr4b mutants in both shape and tumbling behavior. It is likely that this tumbling motion of mutant pllps ensures that, sooner or later, randomly distributed wild-type cells can sample positions at the front edge of the mutant cluster. Once there, the SDF1a-mediated activation of Cxcr4b effectively captures these responsive cells by stimulating their polarized migration (Figure 4B; Movie S11). Therefore, the asymmetric trapping of randomly moving wild-type cells by this extrinsic cue defines what will become the leading edge of the rescued primordium. However, this is only the first step toward rescue, and what ensues is the organization of the SDF1a nonresponsive tissue by this one small group of polarized cells, a process that can take hours (Figures 4C and 4F; Movie S13). Furthermore, once forward migration is established, the timing of neuromast deposition from the trailing edge is also rescued such that the spacing of organs resembles the wild-type situation (Figures 4E and 4G). We conclude that the presence of a small group of SDF1a-responsive cells can rescue many aspects of lateral line morphogenesis in cxcr4b mutants. Discussion The chemokine-receptor pair SDF1/Cxcr4 is receiving growing attention due to its role in directing a number of chemotactic migration events in vivo, both physiological and pathological. Here, we have shown that the pllp can migrate efficiently in both directions along the same endogenous stripe of SDF1a. This suggests that it does not normally move up a concentration gradient, as it would be reluctant to migrate toward lower levels of attractant. If the distribution of guidance cue is indeed uniform, why does the lateral line always migrate in the head-to-tail direction during normal development? One plausible explanation comes from the fact that cells of the primordium are organized in a polarized tissue, with cells closest to the ganglion generating deposited sensory organs. While the primordium requires Cxcr4b to stretch and migrate along the SDF1a source, two pieces of evidence suggest that intrinsic polarity strongly biases the directionality of migration. First, whenever the pllp migrates in the reverse direction in fss mutants, it does so by performing a complicated U-turn maneuver, in which the front regions continue to lead, rather than simply reversing with the former trailing edge guiding. Secondly, when large numbers of wild-type cells were found at the trailing edge of mutant primordia, as was often the case in transplantation experiments, they did not cause a reversal of primordium polarity and backward migration. We therefore conclude that the normal directionality of migration is determined by the organization of the moving tissue itself, rather than by instructive graded cues present in the environment. Through the work presented here, we show that a small number of guidance receptor-expressing cells are able to ensure the directed migration of a large number of mutant neighbors. We think that Cxcr4b-expressing cells do not physically carry mutant neighbors, but rather point these motile cells in the right direction. How does the guiding influence of a small number of cells spread through multicellular tissues? As the pllp is a tightly adherent tissue, it is likely that leader cells guide uninformed neighbors by exerting mechanical force upon them through cell-cell contacts. Alternatively, the leading cells could form a polarized source of a diffusible chemoattractant that directs the migration of followers, as is the case in Dictyostelium slugs, the paradigm of self-organizing migration (Dormann and Weijer, 2001). Single cxcr4b mutant cells within the primordium display dynamic filopodia, indicating that they could indeed be responding to diffusible cues other than the SDF1a stripe. Regardless of what induces this behavior, it suggests that mutant cells, rather than being passive hitchhikers, make an active contribution to lateral line migration. The lateral line primordium is therefore a labile tissue whose shape can be reorganized by SDF1a-responsive cells. Interestingly, mosaic border cell clusters that juxtapose wild-type and immotile mutants give a similar result in which wild-type cells always end up at the front (Rorth et al., 2000). It is tempting to speculate that many other migrating tissues are organized in a similar manner to the lateral line, where guidance is mediated by small numbers of cells whose right to lead is not determined by birth, but rather is earned on the basis of continued guidance receptor activation. Experimental Procedures Zebrafish Strains The following mutant alleles were used in this study: fss te314, ody JIO049.

Cxcr4-Mediated Tissue Migration 679 Generation of the CldnBGFP Transgenic Line Eight kilobases of sequence directly upstream of the Claudin B start codon were amplified from BAC zk241f11 by using the Expand Long Template PCR System (Roche). The resultant fragment was cloned into a vector containing lynegfppa (Koster and Fraser, 2001) flanked by sites for I-SceI, and the resultant construct was injected into one-cell zebrafish embryos by following the meganuclease transgenesis protocol (Thermes et al., 2002). In Situ Hybridization In situ hybridization was carried out with full-length Claudin B and SDF1a probes by following described protocols (Ober and Schulte-Merker, 1999). Mosaic Analysis Genetic mosaics were generated by transplantation by following standard protocols. Briefly, donor embryos were injected at the one-cell stage with 2.5% of the lineage tracer rhodamine dextran (Molecular Probes) and were allowed to develop until the blastula stage. Approximately 20 30 cells were then transplanted into agematched cldnbgfp-positive host embryos. The next day, recipient embryos were screened for the presence of red cells in the pllp by using a Zeiss 510 Meta microscope. To visualize single-cell morphology, cldnbgfp donor cells were transplanted into host embryos that were labeled at the one-cell stage by injection with 0.2 ng mrna encoding mrfp (Gong et al., 2004). Time-Lapse Imaging Embryos were anesthetized in 0.01% tricaine and embedded in 1.5% low-melting point agarose. Time-lapse analysis was carried out on a Zeiss 510 Meta confocal microscope with 103/0.3NA or 203/0.70NA objectives and the 488 nm and 543 nm laserlines. Usually, z-stacks of w10 mm were captured and were then flattened by maximum projection. High-power movies of the wild-type pllp migration were captured by using a Perkin Ultraview LCS spinning disc with a 603/1.4NA Nikon objective. Data Analysis Z-stacks of stage-matched primordia were captured by using identical settings. A region of interest was defined with a box of 150 mm by 25 mm by using Metamorph software (Universal Imaging). Intensity profiles of the red channel were generated by using the Linescan option in Metamorph, and the values were plotted in Excel. Kymographs were generated from time-lapse movies by rotating images by 270º around the y axis with the TransformJ tool of ImageJ (http://rsb.info.nih.gov/ij). Supplemental Data Supplemental Data including Figure S1 and 13 movies are available at http://www.developmentalcell.com/cgi/content/full/10/5/ 673/DC1/. Acknowledgments Part of this work was carried out in the lab of Janni Nüsslein-Volhard (MPI Tübingen). We are grateful for her support and encouragement. We are indebted to Hans-Martin Maischein for expert assistance with zebrafish transplantation and Martina Rembold for providing the mrfp mrna. We thank Jan Huisken and Bernhard Götze for help with image processing, and Nick Foulkes, Virginie Lecaudey, and Francesca Peri for valuable comments on the manuscript. Received: November 24, 2005 Revised: February 10, 2006 Accepted: February 16, 2006 Published: May 8, 2006 References Affolter, M., and Weijer, C.J. (2005). Signaling to cytoskeletal dynamics during chemotaxis. Dev. Cell 9, 19 34. Belmadani, A., Tran, P.B., Ren, D., Assimacopoulos, S., Grove, E.A., and Miller, R.J. (2005). The chemokine stromal cell-derived factor-1 regulates the migration of sensory neuron progenitors. J. Neurosci. 25, 3995 4003. David, N.B., Sapede, D., Saint-Etienne, L., Thisse, C., Thisse, B., Dambly-Chaudiere, C., Rosa, F.M., and Ghysen, A. (2002). Molecular basis of cell migration in the fish lateral line: role of the chemokine receptor CXCR4 and of its ligand, SDF1. Proc. Natl. Acad. Sci. USA 99, 16297 16302. Doitsidou, M., Reichman-Fried, M., Stebler, J., Koprunner, M., Dorries, J., Meyer, D., Esguerra, C.V., Leung, T., and Raz, E. (2002). Guidance of primordial germ cell migration by the chemokine SDF-1. Cell 111, 647 659. Dormann, D., and Weijer, C.J. (2001). Propagating chemoattractant waves coordinate periodic cell movement in Dictyostelium slugs. Development 128, 4535 4543. Friedl, P. (2004). Prespecification and plasticity: shifting mechanisms of cell migration. Curr. Opin. Cell Biol. 16, 14 24. Fulga, T.A., and Rorth, P. (2002). Invasive cell migration is initiated by guided growth of long cellular extensions. Nat. Cell Biol. 4, 715 719. Ghysen, A., and Dambly-Chaudiere, C. (2004). Development of the zebrafish lateral line. Curr. Opin. Neurobiol. 14, 67 73. Gilmour, D., Knaut, H., Maischein, H.M., and Nusslein-Volhard, C. (2004). Towing of sensory axons by their migrating target cells in vivo. Nat. Neurosci. 7, 491 492. Gilmour, D.T., Maischein, H.M., and Nusslein-Volhard, C. (2002). Migration and function of a glial subtype in the vertebrate peripheral nervous system. Neuron 34, 577 588. Gong, Y., Mo, C., and Fraser, S.E. (2004). Planar cell polarity signalling controls cell division orientation during zebrafish gastrulation. Nature 430, 689 693. Kerstetter, A.E., Azodi, E., Marrs, J.A., and Liu, Q. (2004). Cadherin-2 function in the cranial ganglia and lateral line system of developing zebrafish. Dev. Dyn. 230, 137 143. Knaut, H., Werz, C., Geisler, R., and Nusslein-Volhard, C. (2003). A zebrafish homologue of the chemokine receptor Cxcr4 is a germcell guidance receptor. Nature 421, 279 282. Kollmar, R., Nakamura, S.K., Kappler, J.A., and Hudspeth, A.J. (2001). Expression and phylogeny of claudins in vertebrate primordia. Proc. Natl. Acad. Sci. USA 98, 10196 10201. Koster, R.W., and Fraser, S.E. (2001). Tracing transgene expression in living zebrafish embryos. Dev. Biol. 233, 329 346. Li, Q., Shirabe, K., and Kuwada, J.Y. (2004). Chemokine signaling regulates sensory cell migration in zebrafish. Dev. Biol. 269, 123 136. Lopez-Schier, H., Starr, C.J., Kappler, J.A., Kollmar, R., and Hudspeth, A.J. (2004). Directional cell migration establishes the axes of planar polarity in the posterior lateral-line organ of the zebrafish. Dev. Cell 7, 401 412. Mine, N., Iwamoto, R., and Mekada, E. (2005). HB-EGF promotes epithelial cell migration in eyelid development. Development 132, 4317 4326. Nikaido, M., Kawakami, A., Sawada, A., Furutani-Seiki, M., Takeda, H., and Araki, K. (2002). Tbx24, encoding a T-box protein, is mutated in the zebrafish somite-segmentation mutant fused somites. Nat. Genet. 31, 195 199. Ober, E.A., and Schulte-Merker, S. (1999). Signals from the yolk cell induce mesoderm, neuroectoderm, the trunk organizer, and the notochord in zebrafish. Dev. Biol. 215, 167 181. Peled, A., Petit, I., Kollet, O., Magid, M., Ponomaryov, T., Byk, T., Nagler, A., Ben-Hur, H., Many, A., Shultz, L., et al. (1999). Dependence of human stem cell engraftment and repopulation of NOD/SCID mice on CXCR4. Science 283, 845 848. Ribeiro, C., Ebner, A., and Affolter, M. (2002). In vivo imaging reveals different cellular functions for FGF and Dpp signaling in tracheal branching morphogenesis. Dev. Cell 2, 677 683. Rorth, P., Szabo, K., and Texido, G. (2000). The level of C/EBP protein is critical for cell migration during Drosophila oogenesis and is tightly controlled by regulated degradation. Mol. Cell 6, 23 30. Rossi, D., and Zlotnik, A. (2000). The biology of chemokines and their receptors. Annu. Rev. Immunol. 18, 217 242.

Developmental Cell 680 Smith, S.C., Lannoo, M.J., and Armstrong, J.B. (1990). Development of the mechanoreceptive lateral-line system in the axolotl: placode specification, guidance of migration, and the origin of neuromast polarity. Anat. Embryol. (Berl.) 182, 171 180. Stone, L. (1923). On rotation of the lateral line sense organs in Amblystoma punctatum. Abst. Proc. Am. Assoc. Anat. Anat. Rec. 25, 114. Thermes, V., Grabher, C., Ristoratore, F., Bourrat, F., Choulika, A., Wittbrodt, J., and Joly, J.S. (2002). I-SceI meganuclease mediates highly efficient transgenesis in fish. Mech. Dev. 118, 91 98. Tran, P.B., and Miller, R.J. (2003). Chemokine receptors: signposts to brain development and disease. Nat. Rev. Neurosci. 4, 444 455.