Open Access
How to translate text using browser tools
1 December 1998 In vitro Control of Embryonic Axis Formation by Activin A, Concanavalin A, and Retinoic Acid in Xenopus laevis
Naomi Moriya, Chika Yokota, Takashi Ariizumi, Makoto Asashima
Author Affiliations +
Abstract

We have demonstrated that the tissue differentiation patterns along the dorsoventral and anteroposterior axes can be controlled by a combination of activin A, concanavalin A (Con A), and retinoic acid. Xenopus blastula animal caps, normally fated to form epidermal tissues, differentiated into ventral mesoderm tissues such as coelomic epithelium and blood-like cells following treatment with activin A (0.5 ng/ml). Dorsal mesoderm tissues like muscle and notochord, were induced by graded addition of Con A. Conversely, Con A (1 mg/ml) induced anterior neural tissues, forebrain accompanied by eyes and cement glands, in the animal caps. Posterior neural tissues, hindbrain with ear vesicles and spinal cord, were induced by graded addition of activin A. Retinoic acid was also capable of shifting the Con A-induced anterior neural tissues to more posterior tissue phenotypes, however, its caudalizing activity was slightly different from that of activin A. These results suggest that the concentration gradients of these three factors can regulate the differentiation patterns along the embryonic axes. The present study provides a suitable test system for analyzing the establishment of the fundamental body plan in early vertebrate development.

INTRODUCTION

The establishment of the embryonic axes during vertebrate development has been widely studied in amphibians. After the discovery of Spemann's organizer (Spemann and Mangold, 1924), many investigators grappled with analysis of axis formation (reviewed by Sasai and De Robertis, 1997). As a result of these attempts, several models based on concentration gradients of inducer molecules were proposed. In the “two-gradient hypothesis” proposed by Toivonen and Saxén (1955), concentration gradients of two distinct agents, a neuralizing agent (N) and a mesodermalizing agent (M), were postulated to occur in early amphibian embryos. According to this hypothesis, each region of the embryo has a different set of concentrations of the two agents, which leads to the establishment of anteroposterior (A-P) and dorsoventral (D-V) axes. In A-P patterning, anterior neural tissues such as the forebrain, the eye, and the nose are induced by the N agent alone, while varying N/M ratios determine the fate of posterior neural tissues (hindbrain and spinal cord). Another appealing model for A-P specification of the central nervous system (CNS) was proposed by Nieuwkoop and his colleague (1952). In the “twosignal model”, the anterior CNS is induced by the first signal (activator) emanating from the dorsal mesoderm. It may be shifted towards more posterior CNS under the influence of a second signal (transformer) released by the axial mesoderm.

Recently, several factors have been identified as candidates for these agents or signals. Noggin, chordin, and follistatin are the most likely candidates for the N agent or the activator. They are secreted by the dorsal mesoderm before and during gastrulation, and can regulate neural induction and dorsalization of ventral mesoderm explants like Spemann's organizer (Lamb et al., 1993; Hemmati-Brivanlou et al., 1994; Sasai et al., 1995, 1996; Piccolo et al., 1996; Zimmerman et al., 1996). Although concanavalin A (Con A) is not an endogenous factor, it also induces anterior neural tissues in animal cap explants (Takata et al., 1981, 1984; Grunz, 1985) and dorsalizes ventral mesoderm explants in Cynops embryos (Diaz et al., 1990). One of the candidates for the M agent, activin, is known to induce mesodermal tissues along the D-V axis in a dose-dependent manner (Asashima et al., 1990; Green and Smith, 1990; Ariizumi et al., 1991), however, its role in A-P patterning remains unclear. On the other hand, basic fibroblast growth factor (bFGF; Kimelman and Kirschner, 1987; Slack et al., 1987; Cox and Hemmati-Brivanlou, 1995; Kengaku and Okamoto, 1995; Lamb and Harland, 1995), Wnt3A (McGrew et al., 1995) and retinoic acid (RA; Durston et al., 1989; Ruiz i Altaba and Jessell, 1991; Lopez and Carrasco, 1992) are considered candidates for the “transformer”. Recent molecular studies have revealed that these molecules can enhance the expression of posterior markers in anterior tissues. These findings suggest that certain factors regulate the establishment of embryonic axes. However, most of the experiments were carried out using whole embryos or regions that can differentiate by themselves. Further analysis of the patterning of tissue differentiation along the body axes will require a simpler experiment system.

In this study, we employed Xenopus animal caps, which are fated to become atypical epidermis unless influenced by inducing factors. They were treated with various ratios of concentrations of activin A, Con A, and RA to examine whether cell and tissue differentiation along the D-V and A-P axes could be controlled by concentration gradients of these factors in vitro. First, the dorsalizing activity of Con A on the activininduced ventral mesoderm was examined (series I), and second, the caudalizing activities of activin A and RA on the Con A-induced anterior neural tissues were compared (series II and III). The results provide a simple system suitable for analysis of pattern formation along the embryonic axes in vitro.

MATERIALS AND METHODS

Eggs and embryos

Xenopus laevis eggs were obtained by injecting the dorsal lymph sacs of males and females with 600 IU of hCG (GESTRON; Denka Seiyaku, Japan). Late blastulae (stage 9; Nieuwkoop and Faber, 1956) were dejellied with 4.5% cysteine hydrochloride (pH 7.8) dissolved in Steinberg's solution (SS; 58.00 mM NaCl, 0.67 mM KCl, 0.34 mM Ca(NO3)2, 0.83 mM MgSO4, 3.00 mM HEPES and 100 mg/l kanamycin sulfate, pH 7.4). The vitelline membranes were manually removed with a pair of watchmaker's tweezers.

Test solutions

Human recombinant activin A (kindly provided by Dr. Y. Eto, Central Research Laboratories of Ajinomoto Co. Inc., Kawasaki, Japan) and concanavalin A (Con A; L-1104, E-Y LABS, USA) were dissolved in SS. Bovine serum albumin (BSA; A-7888, Sigma, USA) was added to SS to a final concentration of 1 mg/ml to prevent adsorption of activin A and Con A onto the plastic surfaces. Retinoic acid (RA; R2625, Sigma, USA) was dissolved in absolute ethanol at a concentration of 10−2 M. These three solutions were mixed as follows (final concentrations):

i0289-0003-15-6-879-t04.gif

Surgical operation and histology

Animal caps dissected from late blastulae (stage 9) to a size of 0.5 mm × 0.5 mm were treated with the test solutions for 3 hr at 20°C. After two washes with SS, they were cultured for 4 days at 20°C in SS. The animal cap explants were fixed with Bouin's fluid for 3 hr. After dehydration, they were embedded in paraffin and cut into 6 μm sections. The sections were then stained with hematoxylin and eosin.

RT-PCR analysis

Extraction of total RNA, RT-PCR reactions, and subcloning and sequencing of RT-PCR fragments were performed as previously described, with one modification, the RT-PCR reaction was carried out in a 50 μl volume (Yokota et al., 1998). N-CAM (Kintner and Melton, 1987), XAG1 (Sive et al., 1989), Xotx2 (Pannese et al., 1995), En2 (Hemmati-Brivanlou et al., 1991), Krox20 (Bradley et al., 1993), XlHbox 6 (Wright et al., 1990), alpha skeletal muscle actin (muscle-specific actin; ms-actin; Stutz and Spohr, 1986) and EF1-α (Kreig et al., 1989) were analyzed as previously described (Hemmati-Brivanlou and Melton, 1994; Hatada et al., 1995; Lai et al., 1995; Sasai et al., 1995).

RESULTS

Series I: Con A dorsalizes activin A-induced ventral mesoderm

Control animal caps, incubated in SS without activin A and Con A, formed atypical epidermis (Fig. 1A). Ventral mesoderm tissues such as the coelomic epithelium and blood-like cells (77%) were induced by 0.5 ng/ml of activin A (Fig. 1B and Table 1). These tissues were often accompanied by small blocks of muscle cells (80%), and all explants were surrounded by the epidermis (100%). As the concentrations of Con A added to SS plus 0.5 ng/ml activin A increased (50-300 μg/ml), increasingly dorsal types of mesoderm were induced in a concentration-dependent manner. Addition of 50 μg/ml of Con A markedly reduced the frequency of ventral mesoderm (24%). The frequency of muscle remained high (90%), and well-developed muscle cells were seen in the explants (Fig. 1C). The most dorsal mesoderm, the notochord, was also induced at a high frequency (83%). At 100 μg/ml of Con A, the frequency of muscle formation was decreased (60%), while that of notochord formation remained high (87%) (Fig. 1D). At 300 μg/ml of Con A, no muscle was induced, and the frequency of notochord formation was reduced (50%). The frequencies of neural tissue formation were concentration-dependently increased by Con A and reached a plateau (97%) at 100 μg/ml. No endodermal tissues were induced in any specimens.

Fig. 1

Con A dorsalized activin A-induced ventral mesoderm. (A) Control explant cultured without activin A and Con A. (B) Activin A at 0.5 ng/ml induced ventral mesoderm such as coelomic epithelium (Ce). (C) Activin A (0.5 ng/ml) with Con A (50 μg/ml) induced a mass of muscle cells (Mu). (D) Activin A (0.5 ng/ml) with Con A (100 μg/ml) induced notochord (No). Bars: 200 μm.

i0289-0003-15-6-879-f01.gif

Table 1

Differentiation of animal caps treated with activin A plus Con A (series I)

i0289-0003-15-6-879-t01.gif

Activin A at 0.1 ng/ml also induced ventral mesodermal tissues in the animal caps. However, the frequencies were relatively low (30%), and most explants remained undifferentiated (atypical epidermis, 70%). The dorsalizing effects of Con A on the mesoderm could not be seen in these explants, except for a reduction in the frequency of ventral mesoderm. Neural tissues and a cement gland were induced by the addition of Con A (Table 1).

Series II: Activin A caudalizes Con A-induced anterior neural tissues

Neural tissues and a cement gland were induced by 500 μg/ml of Con A at high frequencies (95% and 93%, respectively, Table 2). The neural tissue exhibited a forebrain-like appearance in histological sections (Fig. 2A) and was accompanied by eyes or eye fragments in some specimens (18%). As the concentrations of activin A added to Con A (500 μg/ml) increased (0.5-1 ng/ml), the frequency of cement gland formation decreased, while that of notochord formation increased in a concentration-dependent manner. When activin A (0.5 ng/ml) was added, ear vesicles were induced (11%), but they were not induced by Con A alone. At a high concentration of activin A (1 ng/ml), the eye and eye fragments failed to develop. Although neural tissues were always induced at high frequencies (95% in each combination), the types of neural tissues changed from forebrain-type to spinal cord-type depending on the concentration of activin A (Fig. 2E).

Table 2

Differentiation of animal caps treated with Con A plus activin A (series II)

i0289-0003-15-6-879-t02.gif

Fig. 2

Activin A and RA caudalized Con A-induced anterior neural tissue. (A) Con A (0.5-1 mg/ml) alone induced forebrain (Fb) accompanied by cement gland (Cg). (B) Con A (1 mg/ml) with a low concentration of RA (10−8 M) induced forebrain with an eye (Ey). The explants treated with Con A and a low concentration of activin A (0.5 ng/ml) also differentiated into these structures. (C) Con A (1 mg/ml) with a middle concentration of RA (10−6 M) induced hindbrain (Hb) with ear vesicles (Ev). The explants treated with Con A and higher concentrations of activin A (0.5-1 ng/ml) also differentiated into these structures. (D) A high concentration of RA (10−4 M) with Con A induced nonspecific neural tissue. (E) A high concentration of activin A (1 ng/ml) with Con A induced spinal cord (Sc). Mu: muscle. Bars: 100 μm.

i0289-0003-15-6-879-f02.gif

Series III: RA also caudalizes Con A-induced anterior neural tissues

Neural tissues were induced at every combination of concentrations of Con A and RA (Table 3). Animal caps treated with 1 mg/ml of Con A solely differentiated into the forebrain accompanied by eyes (tapetum with or without lens, 29%) and a cement gland (95%). The addition of a low concentration of RA (10−8 M) to the Con A solution increased the frequencies of formation of eyes or eye fragments (45%) (Fig. 2B). At a higher concentration of RA (10−7 M), differentiation of both the cement gland and the eye were decreased (45% and 0%, respectively). A hindbrain with ear vesicles was induced (38%) when 10−6 M of RA was added to the solution (Fig. 2C). As the RA concentrations increased further (10−5 and 10−4 M), the frequencies of ear vesicle formation gradually diminished. Although neural tissues were frequently induced at higher concentrations of RA, they did not exhibit any of the characteristics found in the spinal cord of normal embryos (Fig. 2D).

Table 3

Differentiation of animal caps treated with Con A plus RA (series III)

i0289-0003-15-6-879-t03.gif

RT-PCR analysis was performed to confirm that RA treatment affected patterns of gene expression in the explants. Con A (1 mg/ml) induced expression of the panneural marker, N-CAM, regardless of the dose of RA added (Fig. 3, lanes 1-4). The anterior markers, XAG1 and Xotx2, were strongly induced by treatment with Con A alone (Fig. 3, lane 1). Expression of these markers was reduced by graded addition of RA, while expression of the posterior neural markers, En2 and Krox20, was enhanced by high concentrations of RA (Fig. 3, lanes 1-4). En2 expression was most strongly induced by 10−6 M of RA (Fig. 3, lane 3). XlHbox 6 and ms-actin, the most posterior neural and mesodermal markers, respectively, were not induced by treatment with Con A and RA (Fig. 3, lanes 1-5). Control explants did not express any markers tested (Fig. 3, lane 6). Although explants treated with RA alone expressed Krox20, the expression was very weak as compared with the explants treated with both Con A and RA (Fig. 3, lanes 4, 5).

Fig. 3

RA caudalized gene expression patterns induced by Con A. Ectoderm at stage 9 was treated with Con A and RA for 3 hr, cultured for 4 days, and subjected to the RT-PCR analysis. Con A alone or with RA at low concentrations induced the anterior markers, XAG1 (cement gland) and Xotx2 (forebrain) (lanes 1-3). The addition of RA reduced the expression of these anterior markers and enhanced posterior markers, En2 (midbrain-hindbrain boundary) and Krox20 (hindbrain rhombomeres 3 and 5) (lanes 1-4). The most posterior marker, XlHbox 6 (spinal cord), and ms-actin (muscle) were not induced by Con A and RA (lanes 1-5). EF1-α RT+ is a loading control and RT− is a negative control. Lanes 1-4: Con A (1 mg/ml) plus RA (0, 10−7, 10−6, 10−5 M). Lane 5: RA (10−5 M) alone. Lane 6: control. Lane 7: whole embryo.

i0289-0003-15-6-879-f03.gif

DISCUSSION

Dorsalizing activity of anterior neural inducers

Con A was originally identified as a neural inducer in isolated Cynops (Takata et al., 1981, 1984) and Xenopus (Grunz, 1985) animal caps. Neural tissues induced in animal caps have characteristics of the forebrain of normal embryos. The same molecule was later found to dorsalize Cynops ventral marginal zone (VMZ) explants. The VMZ explants differentiated into blood cells, pronephros, muscle, and notochord depending on the concentrations of Con A added (Diaz et al., 1990). In the present study, the activin-treated Xenopus animal caps fated to differentiate into ventral mesoderm were also dorsalized by the addition of Con A (series I). Other potent neural inducers, noggin, chordin, and follistatin, induce the expression of panneural marker and anterior neural marker genes in the animal caps (Lamb et al., 1993; Hemmati-Brivanlou et al., 1994; Sasai et al., 1995). Noggin and chordin enhance the expression of muscle actin in Xenopus VMZ explants (Lamb et al., 1993; Piccolo et al., 1996). The VMZ explants isolated from follistatin-mRNA-injected embryos exhibit an elongation movement that mimics the convergent extension of the dorsal mesoderm during gastrulation (Sasai et al., 1995). These findings suggest that the anterior neural inducers may also act as dorsalizing factors on the ventral mesoderm.

The bone morphogenetic protein (BMP) signal ventralizes the mesoderm and inhibits the formation of dorsal mesoderm, Spemann's organizer, in early Xenopus embryos (Dale et al., 1992; Jones et al., 1992; Re'em-Kalma et al., 1995; Schmidt et al., 1995). This signal also acts as an anti-neuralizing agent on the ectoderm (Sasai et al., 1995). The neuralizing/dorsalizing activities of noggin, chordin, and follistatin may result from inhibition of the BMP signal on the ectoderm/mesoderm. Chordin and noggin are known to bind to BMP-4 directly and inhibit the anti-neuralization and ventralization mediated by the BMP signal (Sasai et al., 1995; Piccolo et al., 1996; Zimmerman et al., 1996). Although it is unknown whether Con A can bind to BMP, Con A has been reported to bind to the subunit of the Ca2+ channel and increase intracellular Ca2+ concentration in ectoderm explants. When the increase in Ca2+ was inhibited, neural induction did not occur (Moreau et al., 1994). Thus, there may be a close connection between the concentration of Ca2+ and neural differentiation. Although Con A is not a natural neural inducer, it is a useful reagent for investigating the mechanisms of neuralization and dorsalization. It can induce well-differentiated neural tissue similar to that of the normal embryo, and it possesses neuralizing and dorsalizing activity similar to the natural neural inducers.

Caudalizing activity of activin A and RA

As described in previous reports (Takata et al., 1981, 1984; Grunz, 1985), Con A-treated animal caps differentiated into anterior neural tissues such as the forebrain and the eye. The regional specification of these neural tissues was shifted posteriorly by the addition of activin A (series II). Activin A itself induces a variety of mesodermal tissues along the D-V axis in a concentration-dependent manner (Green and Smith, 1990; Ariizumi et al., 1991). In this study, a low concentration of activin A, insufficient to induce dorsal mesoderm (muscle and notochord) alone, induced these tissues when Con A was added (series I). Takaya (1977) reported that the A-P patterning of the CNS in Cynops embryos is affected by the quality and quantity of surrounding mesodermal tissues. Formation of the forebrain is always accompanied by a small amount of mesenchyme, while the hindbrain and the spinal cord are formed in the presence of the notochord and somitic muscle. In the present study, differentiation of the ear vesicle and the spinal cord was always accompanied by notochord formation. It has been reported that mesodermalized ectoderm induces neural tissues as a secondary induction (Suzuki et al., 1986). It is likely that the caudalization of Con A-induced anterior CNS is caused by the dorsal mesoderm initially induced by activin A and Con A. The signal released by activin A / Con A-induced dorsal mesoderm is unknown at present. However, several candidate caudalizing agents have been reported. bFGF can modify the patterns of gene expression from the anterior to posterior type in animal caps fated to form anterior neural tissues by treatment with noggin or cell-dissociation (Cox and Hemmati-Brivanlou, 1995; Kengaku and Okamoto, 1995; Lamb and Harland, 1995). Wnt3A and β-catenin also enhance the expression of posterior marker genes in anterior neural tissue induced in the animal cap explants (McGrew et al., 1995).

Another candidate caudalizing signal, RA, affects the AP patterning of the CNS. RA and its receptor are localized in the posterior region of Xenopus embryos (Ellinger-Ziegelbauer and Dreyer, 1991; Lopez and Carrasco, 1992; Chen et al., 1994). RA-treated Xenopus embryos have reduced anterior neural structures, as RA transforms the CNS from anterior to the posterior type (Ruiz i Altaba and Jessel, 1991; Lopez and Carrasco, 1992). However, the degree of caudalizing effect of RA is slightly different from that of the mesoderm inducers, as shown above. In the present study, RA led the Con A-induced anterior neural tissue to form hindbrain with ear vesicles. Recently, it was reported that hindbrain is the most sensitive region of the CNS to RA (Kolm et al., 1997; van der Wees et al., 1998). The most posterior CNS region, the spinal cord, was not induced even when the RA concentration was increased up to 10−4 M (series III). Spinal cord formation may require additional factor(s) such as bFGF and Wnt3A. In addition to the caudalizing activity, we have reported a lateralizing activity of RA on activin-induced dorsal mesoderm (Moriya et al., 1993). Therefore, RA may be capable of modifying the differentiation of neural and mesodermal tissues toward the posterior and lateral sides in vitro.

We have demonstrated that three factors, activin A, Con A, and RA, can modify inductive activity of each other on the animal cap cells and that these modifications are concentration-dependent. These results provide a simple and suitable system for analysis of the establishment of the fundamental body plan in vitro.

Acknowledgments

This work was supported by Grants-in-Aid for Scientific Research from the Ministry of Education, Science, Sports and Culture of Japan, and by CREST (Core Research for Evolutional Science and Technology) of the Japan Science and Technology Corporation.

REFERENCES

1.

T. Ariizumi, N. Moriya, H. Uchiyama, and M. Asashima . 1991. Concentration-dependent inducing activity of activin A. Roux's Arch Dev Biol 200:230–233. Google Scholar

2.

M. Asashima, H. Nakano, K. Shimada, K. Kinoshita, K. Ishii, H. Shibai, and N. Ueno . 1990. Mesodermal induction in early amphibian embryos by activin A (erythroid differentiation factor). Roux's Arch Dev Biol 198:330–335. Google Scholar

3.

L. Bradley, A. Snape, S. Bhatt, and D. Wilkinson . 1993. The structure and expression of the Xenopus Krox-20 gene:conserved and divergent patterns of expression in rhombomeres and nerve crest. Mech Dev 40:73–84. Google Scholar

4.

Y. Chen, L. Huang, and M. Solursh . 1994. A concentration gradient of retinoids in the early Xenopus laevis embryo. Dev Biol 161:70–76. Google Scholar

5.

W. G. Cox and A. Hemmati-Brivanlou . 1995. Caudalization of neural fate by tissue recombination and bFGF. Development 121:4349–4358. Google Scholar

6.

L. Dale, G. Howes, B. M. Price, and J. C. Smith . 1992. Bone morphogenetic protein 4:a ventralizing factor in early Xenopus development. Development 115:573–585. Google Scholar

7.

M. R. M. Diaz, T. C. Takahashi, K. Takeshima, and K. Takata . 1990. Concanavalin A acts a factor in establishing the dorso-ventral gradient in the ventral mesoderm of newt gastrula embryos. Develop Growth Differ 32:117–123. Google Scholar

8.

A. J. Durston, J. P. M. Timmermans, W. J. Hage, H. F. J. Hendriks, N. J. de Vries, M. Heideveld, and P. D. Nieuwkoop . 1989. Retinoic acid causes an anteroposterior transformation in the developing central nervous system. Nature 340:140–144. Google Scholar

9.

H. Ellinger-Ziegelbauer and C. Dreyer . 1991. A retinoic acid receptor expressed in the early development of Xenopus laevis. Genes Dev 5:94–104. Google Scholar

10.

J. B. A. Green and J. C. Smith . 1990. Graded changes in dose of a Xenopus activin A homologue elicit stepwise transitions in embryonic cell fate. Nature 347:391–394. Google Scholar

11.

H. Grunz 1985. Effect of Concanavalin A and vegetalizing factor an the outer and inner ectoderm layers of early gastrulae of Xenopus laevis after treatment with cytochalacin B. Cell Differ 16:83–92. Google Scholar

12.

S. Hatada, M. Kinoshita, M. Noda, and M. Asashima . 1995. Identification of a Xenopus glutamine synthetase gene abundantly expressed in the embryonic nervous system but not in adult brain. FEBS Lett 371:287–292. Google Scholar

13.

A. Hemmati-Brivanlou, J. R. de la Torre, C. Holt, and R. M. Harland . 1991. Cephalic expression and molecular characterization of Xenopus En-2. Development 111:715–724. Google Scholar

14.

A. Hemmati-Brivanlou, O. G. Kelley, and D. A. Melton . 1994. Follistatin, an antagonist of activin is expressed in the Spemann organizer and displays direct neuralizing activity. Cell 77:283–295. Google Scholar

15.

A. Hemmati-Brivanlou and D. A. Melton . 1994. Inhibition of activin receptor signaling promotes neuralization in Xenopus. Cell 77:273–281. Google Scholar

16.

C. M. Jones, K. M. Lyons, P. M. Lapan, C. V. Wright, and B. L. Hogan . 1992. DVR-4 (bone morphogenetic protein-4) as a posterior-ventralizing factor in Xenopus mesoderm induction. Development 115:639–647. Google Scholar

17.

M. Kengaku and H. Okamoto . 1995. bFGF as a possible morphogen for the anteroposterior axis of the central nervous system in Xenopus. Development 121:3121–3130. Google Scholar

18.

D. Kimelman and M. Kirschner . 1987. Synergistic induction of mesoderm by FGF and TGF-β and the identification of an mRNA coding for FGF in the early Xenopus embryo. Cell 51:869–877. Google Scholar

19.

C. R. Kintner and D. A. Melton . 1987. Expression of Xenopus N-CAM RNA in ectoderm is an early response to neural induction. Development 99:311–325. Google Scholar

20.

P. J. Kolm, V. Apekin, and H. Sive . 1997. Xenopus hindbrain pattering requires retinoid signaling. Dev Biol 192:1–16. Google Scholar

21.

P. A. Kreig, S. Varnam, M. Wormington, and D. A. Melton . 1989. The mRNA encoding elongation factor 1α (EF-1α) is a major transcript at the mid-blastula transition in Xenopus. Dev Biol 133:93–100. Google Scholar

22.

C-J. Lai, S. C. Ekker, P. A. Beachy, and R. T. Moon . 1995. Patterning of the neural ectoderm of Xenopus laevis by the amino-terminal product of hedgehog autoproteolytic cleavage. Development 121:2349–2360. Google Scholar

23.

T. M. Lamb, A. K. Knecht, W. C. Smith, S. E. Stachel, A. N. Economides, N. Stahl, G. D. Yancopolous, and R. M. Harland . 1993. Neural induction by the secreted polypeptide noggin. Science 262:713–718. Google Scholar

24.

T. M. Lamb and R. M. Harland . 1995. Fibroblast growth factor is a direct neural inducer, which combined with noggin generates anterior-posterior neural pattern. Development 121:3627–3636. Google Scholar

25.

S. L. Lopez and A. E. Carrasco . 1992. Retinoic acid modifies the pattern of cell differentiation in the central nervous system of neurula stage Xenopus embryo. Development 112:945–958. Google Scholar

26.

L. L. McGrew, C-J. Lai, and R. T. Moon . 1995. Specification of the anteroposterior neural axis through synergistic interaction of the Wnt signaling cascade with noggin and follistatin. Dev Biol 172:337–342. Google Scholar

27.

M. Moreau, C. Leclerc, L. Gualandris-Parisot, and A. M. Duprat . 1994. Increased internal Ca2+ mediates neural induction in the amphibian embryo. Proc Natl Acad Sci USA 91:12639–12643. Google Scholar

28.

N. Moriya, H. Uchiyama, and M. Asashima . 1993. Induction of pronephric tubules by activin and retinoic acid in presumptive ectoderm of Xenopus laevis. Develop Growth Differ 35:123–128. Google Scholar

29.

P. D. Nieuwkoop, E. C. Boterenbrood, A. Kremer, F. F. S. N. Bloesma, E. L. M. J. Hoessels, G. Meyer, and F. J. Verheyen . 1952. Activation and organization of the central nervous system in amphibians. J Exp Zool 120:1–108. Google Scholar

30.

P. D. Nieuwkoop and J. Faber . 1956. Normal Table of Xenopus laevis (Daudin). North-Holland Publ. Amsterdam. Google Scholar

31.

M. Pannese, C. Polo, M. Andreazzoli, R. Vignali, B. Kablar, G. Barasacchi, E-J. Dekker, and E. Boncinelli . 1995. The Xenopus homologue of Otx2 is a maternal homeobox gene that demarcates and specifies anterior body regions. Development 121:707–720. Google Scholar

32.

S. Piccolo, Y. Sasai, B. Lu, and E. M. De Robertis . 1996. Dorsoventral patterning in Xenopus:inhibition of ventral signal by direct binding of chordin to BMP-4. Cell 86:589–598. Google Scholar

33.

Y. Re'em-Kalma, T. Lamb, and D. Frank . 1995. Competition between noggin and bone morphogenetic protein 4 activities may regulate dorsalization during Xenopus development. Proc Natl Acad Sci USA 92:12141–12145. Google Scholar

34.

A. Ruiz i Altaba and T. Jessell . 1991. Retinoic acid modifies the pattern of cell differentiation in the central nervous system of neurula stage Xenopus embryos. Development 112:945–958. Google Scholar

35.

Y. Sasai, B. Lu, H. Steinbeisser, and E. M. De Robertis . 1995. Regulation of neural induction by the Chd and Bmp-4 antagonistic patterning signals in Xenopus. Nature 376:333–336. Google Scholar

36.

Y. Sasai, B. Lu, S. Piccolo, and E. M. De Robertis . 1996. Endoderm induction by the organizer-secreted factors chordin and noggin in Xenopus animal caps. EMBO J 15:4547–4555. Google Scholar

37.

Y. Sasai and E. M. De Robertis . 1997. Ectodermal patterning in vertebrate embryos. Dev Biol 182:5–20. Google Scholar

38.

J. E. Schmidt, A. Suzuki, N. Ueno, and D. Kimelman . 1995. Localized BMP-4 mediates dorsal/ventral patterning in the early Xenopus embryo. Dev Biol 169:37–50. Google Scholar

39.

H. L. Sive, K. Hattori, and H. Weintraub . 1989. Progressive determination during formation of the anteroposterior axis in Xenopus laevis. Cell 58:171–180. Google Scholar

40.

J. M. W. Slack, B. G. Darligton, J. K. Heath, and S. F. Godsave . 1987. Mesoderm induction in early Xenopus embryos by heparin-binding growth factors. Nature 326:197–200. Google Scholar

41.

H. Spemann and H. Mangold . 1924. Über Induktion von Embryonalanlagen durch Implantation artfremder Organisatoren. Arch mikrosk Anat EntwMech 100:599–638. Google Scholar

42.

F. Stutz and G. Spohr . 1986. Isolation and characterization of sarcomic actin genes expressed in Xenopus laevis embryos. J Mol Biol 187:349–361. Google Scholar

43.

A. S. Suzuki, Y. Yoshimura, and Y. Yano . 1986. Neural-inducing activity of newly mesodermalized ectoderm. Roux's Arch Dev Biol 195:168–172. Google Scholar

44.

K. Takata, K. Y. Yamamoto, and R. Ozawa . 1981. Use of lectins as probes for analyzing embryonic induction. Roux's Arch Dev Biol 190:92–96. Google Scholar

45.

K. Takata, K. Y. Yamamoto, I. Ishii, and N. Takahashi . 1984. Glycoproteins responsive to the neural-inducing effect of Concanavalin A in Cynops presumptive ectoderm. Cell Differ 14:25–31. Google Scholar

46.

H. Takaya 1977. Types of neural tissues induced through the presumptive notochord of newt embryo. Differentiation 7:187–192. Google Scholar

47.

S. Toivonen and L. Saxén . 1955. Über die Induktion des Neural-rohrs bei Triturus Keimen als simultane Leistung des Leber-und Knochenmarkgewebes vom Meerschweinchen. Ann Acad Sci Fenn A IV 30:1–29. Google Scholar

48.

J. van der Wees, J. G. Schilthuis, C. H. Koster, H. Diesveld-Schipper, G. E. Folkers, P. T. van der Saag, M. I. Dawson, K. Shudo, B. van der Burg, and A. J. Durston . 1998. Inhibition of retinoic acid receptor-mediated signaling alters positional identity in the developing hind-brain. Development 125:545–556. Google Scholar

49.

C. V. E. Wright, E. A. Morita, D. J. Wilkin, and E. M. De Robertis . 1990. The Xenopus XlHbox 6 homeo protein, a marker of posterior neural induction, is expressed in proliferating neurons. Development 109:225–234. Google Scholar

50.

C. Yokota, S. Takahashi, A. Eisaki, M. Asashima, S. Akhter, T. Muramatsu, and K. Kadomatsu . 1998. Midkine counteracts the activin signal in mesoderm induction and promotes neural formation. J Biochem 123:339–346. Google Scholar

51.

L. B. Zimmerman, J. M. De Jesus-Escobar, and R. M. Harland . 1996. The Spemann organizer signal noggin binds and inactivate bone morphogenetic protein 4. Cell 86:599–606. Google Scholar
Naomi Moriya, Chika Yokota, Takashi Ariizumi, and Makoto Asashima "In vitro Control of Embryonic Axis Formation by Activin A, Concanavalin A, and Retinoic Acid in Xenopus laevis," Zoological Science 15(6), 879-886, (1 December 1998). https://doi.org/10.2108/zsj.15.879
Received: 7 July 1998; Accepted: 1 September 1998; Published: 1 December 1998
Back to Top