Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Lrp4 in astrocytes modulates glutamatergic transmission

Abstract

Neurotransmission requires precise control of neurotransmitter release from axon terminals. This process is regulated by glial cells; however, the underlying mechanisms are not fully understood. We found that glutamate release in the brain was impaired in mice lacking low-density lipoprotein receptor–related protein 4 (Lrp4), a protein that is critical for neuromuscular junction formation. Electrophysiological studies revealed compromised release probability in astrocyte-specific Lrp4 knockout mice. Lrp4 mutant astrocytes suppressed glutamatergic transmission by enhancing the release of ATP, whose level was elevated in the hippocampus of Lrp4 mutant mice. Consequently, the mutant mice were impaired in locomotor activity and spatial memory and were resistant to seizure induction. These impairments could be ameliorated by blocking the adenosine A1 receptor. The results reveal a critical role for Lrp4, in response to agrin, in modulating astrocytic ATP release and synaptic transmission. Our findings provide insight into the interaction between neurons and astrocytes for synaptic homeostasis and/or plasticity.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Reduced EPSC frequency and impaired synaptic plasticity in GFAP-Lrp4−/− mice.
Figure 2: Reduced excitatory vesicle release probability in the absence of Lrp4.
Figure 3: Astrocytic Lrp4 is critical for glutamatergic transmission.
Figure 4: Astrocytic Lrp4 affects glutamatergic transmission in a non-contact way.
Figure 5: Modulation of astrocytic ATP release by Lrp4.
Figure 6: Agrin regulates astrocytic ATP release.
Figure 7: Ablation of Lrp4 caused abnormal behavior.

Similar content being viewed by others

References

  1. Haydon, P.G. GLIA: listening and talking to the synapse. Nat. Rev. Neurosci. 2, 185–193 (2001).

    Article  CAS  PubMed  Google Scholar 

  2. Perea, G., Navarrete, M. & Araque, A. Tripartite synapses: astrocytes process and control synaptic information. Trends Neurosci. 32, 421–431 (2009).

    Article  CAS  PubMed  Google Scholar 

  3. Clarke, L.E. & Barres, B.A. Emerging roles of astrocytes in neural circuit development. Nat. Rev. Neurosci. 14, 311–321 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Halassa, M.M. & Haydon, P.G. Integrated brain circuits: astrocytic networks modulate neuronal activity and behavior. Annu. Rev. Physiol. 72, 335–355 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Azevedo, F.A. et al. Equal numbers of neuronal and nonneuronal cells make the human brain an isometrically scaled-up primate brain. J. Comp. Neurol. 513, 532–541 (2009).

    Article  PubMed  Google Scholar 

  6. Witcher, M.R., Kirov, S.A. & Harris, K.M. Plasticity of perisynaptic astroglia during synaptogenesis in the mature rat hippocampus. Glia 55, 13–23 (2007).

    Article  PubMed  Google Scholar 

  7. Hamilton, N.B. & Attwell, D. Do astrocytes really exocytose neurotransmitters? Nat. Rev. Neurosci. 11, 227–238 (2010).

    Article  CAS  PubMed  Google Scholar 

  8. Araque, A. et al. Gliotransmitters travel in time and space. Neuron 81, 728–739 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Zhang, J.M. et al. ATP released by astrocytes mediates glutamatergic activity-dependent heterosynaptic suppression. Neuron 40, 971–982 (2003).

    Article  CAS  PubMed  Google Scholar 

  10. Pascual, O. et al. Astrocytic purinergic signaling coordinates synaptic networks. Science 310, 113–116 (2005).

    Article  CAS  PubMed  Google Scholar 

  11. Herz, J. & Strickland, D.K. LRP: a multifunctional scavenger and signaling receptor. J. Clin. Invest. 108, 779–784 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Shen, C., Xiong, W.C. & Mei, L. LRP4 in neuromuscular junction and bone development and diseases. Bone 80, 101–108 (2015).

    Article  CAS  PubMed  Google Scholar 

  13. Kim, N. et al. Lrp4 is a receptor for Agrin and forms a complex with MuSK. Cell 135, 334–342 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Zhang, B. et al. LRP4 serves as a coreceptor of agrin. Neuron 60, 285–297 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Gautam, M. et al. Defective neuromuscular synaptogenesis in agrin-deficient mutant mice. Cell 85, 525–535 (1996).

    Article  CAS  PubMed  Google Scholar 

  16. Weatherbee, S.D., Anderson, K.V. & Niswander, L.A. LDL-receptor-related protein 4 is crucial for formation of the neuromuscular junction. Development 133, 4993–5000 (2006).

    Article  CAS  PubMed  Google Scholar 

  17. Barik, A., Lu, Y. & Sathyamurthy, A. LRP4 is critical for neuromuscular junction maintenance. J. Neurosci. 34, 13892–13905 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  18. Wu, H., Xiong, W.C. & Mei, L. To build a synapse: signaling pathways in neuromuscular junction assembly. Development 137, 1017–1033 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Zong, Y. et al. Structural basis of agrin-LRP4-MuSK signaling. Genes Dev. 26, 247–258 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. DeChiara, T.M. et al. The receptor tyrosine kinase MuSK is required for neuromuscular junction formation in vivo. Cell 85, 501–512 (1996).

    Article  CAS  PubMed  Google Scholar 

  21. Yumoto, N., Kim, N. & Burden, S.J. Lrp4 is a retrograde signal for presynaptic differentiation at neuromuscular synapses. Nature 489, 438–442 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Wu, H. et al. Distinct roles of muscle and motoneuron LRP4 in neuromuscular junction formation. Neuron 75, 94–107 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Tian, Q.B. et al. Interaction of LDL receptor-related protein 4 (LRP4) with postsynaptic scaffold proteins via its C-terminal PDZ domain-binding motif, and its regulation by Ca/calmodulin-dependent protein kinase II. Eur. J. Neurosci. 23, 2864–2876 (2006).

    Article  PubMed  Google Scholar 

  24. Gomez, A.M., Froemke, R.C. & Burden, S.J. Synaptic plasticity and cognitive function are disrupted in the absence of Lrp4. eLife 3, e04287 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  25. Pohlkamp, T. et al. Lrp4 domains differentially regulate limb/brain development and synaptic plasticity. PLoS One 10, e0116701 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Zhuo, L. et al. hGFAP-cre transgenic mice for manipulation of glial and neuronal function in vivo. Genesis 31, 85–94 (2001).

    Article  CAS  PubMed  Google Scholar 

  27. Zucker, R.S. & Regehr, W.G. Short-term synaptic plasticity. Annu. Rev. Physiol. 64, 355–405 (2002).

    Article  CAS  PubMed  Google Scholar 

  28. Rosenmund, C., Clements, J.D. & Westbrook, G.L. Nonuniform probability of glutamate release at a hippocampal synapse. Science 262, 754–757 (1993).

    Article  CAS  PubMed  Google Scholar 

  29. Hessler, N.A., Shirke, A.M. & Malinow, R. The probability of transmitter release at a mammalian central synapse. Nature 366, 569–572 (1993).

    Article  CAS  PubMed  Google Scholar 

  30. Stevens, C.F. & Wang, Y. Facilitation and depression at single central synapses. Neuron 14, 795–802 (1995).

    Article  CAS  PubMed  Google Scholar 

  31. Casper, K.B., Jones, K. & McCarthy, K.D. Characterization of astrocyte-specific conditional knockouts. Genesis 45, 292–299 (2007).

    Article  CAS  PubMed  Google Scholar 

  32. Ge, W.P., Miyawaki, A., Gage, F.H., Jan, Y.N. & Jan, L.Y. Local generation of glia is a major astrocyte source in postnatal cortex. Nature 484, 376–380 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Panatier, A. et al. Glia-derived D-serine controls NMDA receptor activity and synaptic memory. Cell 125, 775–784 (2006).

    Article  CAS  PubMed  Google Scholar 

  34. Yang, Y. et al. Contribution of astrocytes to hippocampal long-term potentiation through release of D-serine. Proc. Natl. Acad. Sci. USA 100, 15194–15199 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Chu, S., Xiong, W. & Parkinson, F.E. Effect of ecto-5′-nucleotidase (eN) in astrocytes on adenosine and inosine formation. Purinergic Signal. 10, 603–609 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Bezakova, G. & Ruegg, M.A. New insights into the roles of agrin. Nat. Rev. Mol. Cell Biol. 4, 295–308 (2003).

    Article  CAS  PubMed  Google Scholar 

  37. Ksiazek, I. et al. Synapse loss in cortex of agrin-deficient mice after genetic rescue of perinatal death. J. Neurosci. 27, 7183–7195 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Garcia-Osta, A. et al. MuSK expressed in the brain mediates cholinergic responses, synaptic plasticity and memory formation. J. Neurosci. 26, 7919–7932 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Morris, R.G., Garrud, P., Rawlins, J.N. & O'Keefe, J. Place navigation impaired in rats with hippocampal lesions. Nature 297, 681–683 (1982).

    Article  CAS  PubMed  Google Scholar 

  40. Curia, G., Longo, D., Biagini, G., Jones, R.S. & Avoli, M. The pilocarpine model of temporal lobe epilepsy. J. Neurosci. Methods 172, 143–157 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Suzuki, T. et al. Efhc1 deficiency causes spontaneous myoclonus and increased seizure susceptibility. Hum. Mol. Genet. 18, 1099–1109 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Hines, D.J. & Haydon, P.G. Astrocytic adenosine: from synapses to psychiatric disorders. Phil. Trans. R. Soc. Lond. B 369, 20130594 (2014).

    Article  CAS  Google Scholar 

  43. Jourdain, P. et al. Glutamate exocytosis from astrocytes controls synaptic strength. Nat. Neurosci. 10, 331–339 (2007).

    Article  CAS  PubMed  Google Scholar 

  44. Li, Y., Krupa, B., Kang, J.S., Bolshakov, V.Y. & Liu, G. Glycine site of NMDA receptor serves as a spatiotemporal detector of synaptic activity patterns. J. Neurophysiol. 102, 578–589 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Coco, S. et al. Storage and release of ATP from astrocytes in culture. J. Biol. Chem. 278, 1354–1362 (2003).

    Article  CAS  PubMed  Google Scholar 

  46. Stout, C.E., Costantin, J.L., Naus, C.C. & Charles, A.C. Intercellular calcium signaling in astrocytes via ATP release through connexin hemichannels. J. Biol. Chem. 277, 10482–10488 (2002).

    Article  CAS  PubMed  Google Scholar 

  47. Suadicani, S.O., Brosnan, C.F. & Scemes, E. P2X7 receptors mediate ATP release and amplification of astrocytic intercellular Ca2+ signaling. J. Neurosci. 26, 1378–1385 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Roman, R.M. et al. Hepatocellular ATP-binding cassette protein expression enhances ATP release and autocrine regulation of cell volume. J. Biol. Chem. 272, 21970–21976 (1997).

    Article  CAS  PubMed  Google Scholar 

  49. Zhang, Z. et al. Regulated ATP release from astrocytes through lysosome exocytosis. Nat. Cell Biol. 9, 945–953 (2007).

    Article  CAS  PubMed  Google Scholar 

  50. Christopherson, K.S. et al. Thrombospondins are astrocyte-secreted proteins that promote CNS synaptogenesis. Cell 120, 421–433 (2005).

    Article  CAS  PubMed  Google Scholar 

  51. Tsien, J.Z. et al. Subregion- and cell type–restricted gene knockout in mouse brain. Cell 87, 1317–1326 (1996).

    Article  CAS  PubMed  Google Scholar 

  52. Shen, C. et al. Antibodies against low-density lipoprotein receptor-related protein 4 induce myasthenia gravis. J. Clin. Invest. 123, 5190–5202 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Ting, A.K. et al. Neuregulin 1 promotes excitatory synapse development and function in GABAergic interneurons. J. Neurosci. 31, 15–25 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Cao, X. et al. Astrocyte-derived ATP modulates depressive-like behaviors. Nat. Med. 19, 773–777 (2013).

    Article  CAS  PubMed  Google Scholar 

  55. Chen, Y.J. et al. ErbB4 in parvalbumin-positive interneurons is critical for neuregulin 1 regulation of long-term potentiation. Proc. Natl. Acad. Sci. USA 107, 21818–21823 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Gil, Z., Connors, B.W. & Amitai, Y. Efficacy of thalamocortical and intracortical synaptic connections: quanta, innervation, and reliability. Neuron 23, 385–397 (1999).

    Article  CAS  PubMed  Google Scholar 

  57. Yin, D.M. et al. Regulation of spine formation by ErbB4 in PV-positive interneurons. J. Neurosci. 33, 19295–19303 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Wen, L. et al. Neuregulin 1 regulates pyramidal neuron activity via ErbB4 in parvalbumin-positive interneurons. Proc. Natl. Acad. Sci. USA 107, 1211–1216 (2010).

    Article  CAS  PubMed  Google Scholar 

  59. Bi, L.L. et al. Amygdala NRG1-ErbB4 is critical for the modulation of anxiety-like behaviors. Neuropsychopharmacology 40, 974–986 (2015).

    Article  CAS  PubMed  Google Scholar 

  60. Yin, D.M. et al. Reversal of behavioral deficits and synaptic dysfunction in mice overexpressing neuregulin 1. Neuron 78, 644–657 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Racine, R.J. Modification of seizure activity by electrical stimulation. II. Motor seizure. Electroencephalogr. Clin. Neurophysiol. 32, 281–294 (1972).

    Article  CAS  PubMed  Google Scholar 

  62. Lu, Y. et al. Maintenance of GABAergic activity by neuregulin 1-ErbB4 in amygdala for fear memory. Neuron 84, 835–846 (2014).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We are grateful to W.-P. Ge (UT Southwestern Medical Center) and K.D. McCarthy (University of North Carolina) for GFAP::CreER mice. We thank members of the Mei and Xiong laboratories for helpful discussions. This work was supported in part by grants from the US National Institutes of Health (L.M., W.-C.X.) and Veterans Affairs (L.M., W.-C.X.), “Thousand Talents” Innovation Project from Jiangxi Province (L.M.), National Natural Science Foundation of China (NSFC, 81471116, B.-M.L), NSFC (81329003; U1201225; 31430032, T.-M.G.), Guangzhou Science and Technology Project (201300000093, T.-M.G.) and Specialized Research Fund for the Doctoral Program of Higher Education of China (20134433130002, T.-M.G.).

Author information

Authors and Affiliations

Authors

Contributions

X.-D.S. and L.M. conceived, designed and directed the project, and wrote the manuscript. X.-D.S performed electrophysiological recordings and analysis. L.L. conducted immunoblots, quantitative reverse-transcription PCR (qRT-PCR) and co-immunoporecipitation. F.L., R.B., A.B. and A.S. conducted Golgi staining, X-gal staining, immunofluorescence staining and analysis. Z.-H.H. conducted immunoblots and astrocyte culture experiments and analysis. J.C.B. performed behavioral tests and microdialysis analysis, with the assistance of Y.-J.C. and D.-M.Y. H.-F.J., S.-M.K. and Y.T. conducted cell culture experiments and analysis. H.W. and C.S., provided and assisted with characterization of Lrp4 mutant mice. T.W.L. conducted spine and synapse analysis. L.X., H.-P.L., J.-X.H. assisted with breeding and genotyping Lrp4 mutant lines. B.-M.L., T.-M.G. and W.-C.X. helped with data interpretation and provided instruction. L.M. supervised the project.

Corresponding authors

Correspondence to Xiang-Dong Sun or Lin Mei.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Hippocampal Lrp4 expression and morphological characterization of GFAP-Lrp4–/– hippocampus and cortex.

(a) Lrp4 expression in the hippocampus at different stages. Shown were representative blots of two independent experiments with similar results. Full-length blots/gels are presented in Supplementary Figure 12.

(b) Quantitative analysis of Lrp4 expression in a. n =3 mice. Lrp4 band density was normalized by the loading control β-actin; values of postnatal day 0 in (a) were taken as 100%.

(c) Cortical sections of control (top) and GFAP-Lrp4–/– (bottom) mice were stained for neuron marker NeuN. Enlarged images of areas in dotted square were shown on the right. Scale bar: left, 250 μm; right, 50 μm.

(d) Quantification of NeuN+ cells in different cortical layers. n = 8 slices of 3 mice for both genotypes. Two-way ANOVA, F(1,70) = 3.736, p = 0.0573.

(e) Dorsal hippocampal sections of control (top) and GFAP-Lrp4–/– (bottom) mice were stained for neuron marker NeuN. Images in dotted areas (marked as a’ in CA1 and a’’ in CA3) were enlarged and shown on the right. Scale bar: left, 200 μm, right, 20 μm.

(f) Quantification of NeuN+ cells in CA1 and CA3 regions in dorsal hippocampus. n = 10 slices, 3 mice for both genotypes. Two-way ANOVA, F(1,36) = 0.1503, p = 0.7005.

Data in (b) were presented as mean ± s.e.m; data in (d, f) were presented as median with interquartile range, whiskers are the minimum and maximum; p > 0.05.

Supplementary Figure 2 Synaptic protein expression and dendritic morphology of CA1 pyramidal neurons in GFAP-Lrp4–/– mice.

(a) Hippocampal tissues were collected from one-month-old control and GFAP-Lrp4–/– mice and homogenized for western blotting. Shown were representative blots of three independent experiments with similar results. Full-length blots/gels are presented in Supplementary Figure 12.

(b) Quantitative analysis of data in (a). n =5 mice of each genotype for Lrp4, Synaptophysin, GluR1, GluR2 and GABAα1; n = 4 mice of each genotype for PSD95 and Gephyin; n = 3 mice of each genotype for GluN1, GluN2A and GluN2B. Band densities of interested proteins were normalized by the loading control β-actin; values of control mice were taken as 100%. Paired student’s t test; Lrp4, t(4) = 28.05, p < 0.0001; PSD95, t(3) = 0.4635, p = 0.6746; Gephyrin, t(3) = 1.54, p = 0.2212; Synaptophysin, t(4) = 0.3952, p = 0.7128; GluR1, t(4) = 0.2008, p = 0.8507; GluR2, t(4) = 0.8446, p = 0.4459; GluN1, t(2) = 0.003932, p = 0.9972; GluN2A, t(2) = 0.3696, p = 0.7472; GluN2B, t(2) = 0.6231, p = 0.5968; GABARα1, t(4) = 2.008, p = 0.1151.

(c) Representative traced CA1 pyramidal neurons were superimposed with concentric circles in Sholl analysis. The radius interval between circles was 20 µm per step, ranging from 10 µm to 290 µm from the center of the neuronal soma.

(d) Quantitative analysis of data in (c). n = 15 neurons from 3 mice per genotype. Two-way ANOVA, F(1,363) = 0.1266, p = 0.7222..

Data in (b) were presented as median with interquartile range, whiskers are the minimum and maximum; data in (d) were presented as mean ± s.e.m; p > 0.05.

Supplementary Figure 3 Electrophysiological characterization of CA1 pyramidal neurons from GFAP-Lrp4–/– mice.

(a) Representative traces of spikes in CA1 pyramidal neurons evoked by injecting depolarizing currents of 50 pA. Scale bar: 100 ms, 20 mV.

(b) Firing rate plotted against increasing injected currents. n = 9 neurons, 3 control mice; n = 8 neurons, 3 GFAP-Lrp4–/– mice. Two-way ANOVA, F(1,150) = 0.006, p = 0.9396.

(c, d) Spike threshold evoked by 50-pA currents and input resistance. n = 9 neurons, 3 control mice; n = 8 neurons, 3 GFAP-Lrp4–/– mice. For (c), student’s t test; t(15) = 0.2535, p = 0.8033; for (d), student’s t test, t(15) = 0.7621, p = 0.4578.

(e) Representative traces of EPSCs evoked at holding potentials (–60 mV to + 60 mV at a step of 20 mV). Scale bar: 20 ms, 60 pA.

(f) Current-voltage plot of peak currents mediated by AMPAR. n = 11 neurons of 3 mice for both genotypes. Two-way ANOVA, F(1,140) = 1.449, p = 0.2307.

(g) Current-voltage plot of NMDAR EPSCs measured 50 ms after the peak of AMPAR EPSCs. n = 11 neurons of 3 mice for both genotypes. Two-way ANOVA, F(1,140) = 0.022, p = 0.8825.

(h) AMPA/NMDA ratio. The ratio was calculated by dividing the amplitude of AMPAR currents measured at –60 mV by the amplitude of NMDAR currents measured 50 ms after the peak at + 40 mV. n = 11 neurons of 3 mice for both genotypes. Student’s t test, t(20) = 0.4578, p = 0.652.

Data in (b, f and g) were presented as mean ± s.e.m; data in (c, d and h) were presented as median with interquartile range, whiskers are the minimum and maximum; p > 0.05.

Supplementary Figure 4 Characterization of spines of CA1 pyramidal neurons of GFAP-Lrp4–/– mice.

(a) Representative Golgi staining images of primary and second/tertiary (sec/ter) dendrites of CA 1 pyramidal neurons of control and GFAP-Lrp4–/– mice. Scale bar, 10 μm.

(b) Quantitative analysis of data in (a). For primary dendrites, n = 46 segments, 46 neurons from 3 mice per genotype, student’s t test, t(90) = 2.055, p = 0.0427; For sec/ter dendrites, n = 80 segments, 46 neurons from 3 control mice, n = 76 segments, 42 neurons from GFAP-Lrp4–/– mice, student’s t test, t(154) = 3.195, p = 0.0017.

(c) Representative EM images of CA1 synapses from control and GFAP-Lrp4–/– mice. Arrows, PSD; arrowheads, vesicles. Scale bar: 200 nm.

(d) Quantitative analysis of PSD length. n = 63 synapses, 3 control mice; n = 67 synapses, 3 GFAP-Lrp4–/– mice. Student’s t test, t(128) = 1.28, p = 0.2029.

(e) Quantitative analysis of synaptic vesicle density. n = 61 synapses, 3 control mice; n = 64 synapses, 3 GFAP-Lrp4–/– mice. Student’s t test, t(123) = 0.3366, p = 0.737.

(f) Quantitative analysis of vesicle diameter. n = 249 vesicles, 3 control mice; n = 263 vesicles, 3 GFAP-Lrp4–/– mice. Student’s t test, t(510) = 0.7743, p = 0.4391.

(g) Quantitative analysis of astrocyte-synaptic membrane distance. n = 164, 3 control mice; n = 172, 3 GFAP-Lrp4–/– mice. Student’s t test, t(334) = 1.339, p = 0.1815.

Data in (b, d, e, f and g) were presented as median with interquartile range, whiskers are the minimum and maximum. *, p < 0.05; **, p < 0.01.

Supplementary Figure 5 Lrp4 expression in astrocytes and characterization of GFAP::CreER mice.

(a) Lrp4 was detectable in homogenate of control, but not GFAP-Lrp4–/–, astrocytes in culture. Astrocytes were isolated from P2-3 mice, cultured in vitro for 7 days and lysed for western analysis. Shown were representative blots of more than three independent experiments with similar results. Full-length blots/gels are presented in Supplementary Figure 12.

(b) Top, GFAP::CreER mice were crossed with reporter mice (tdTomato) to generate GFAP::CreER;tdTomato mice. Bottom, GFAP::CreER;tdTomato at P30 were administered by gavage one time with tamoxifen (100 mg/kg) and sacrificed at P35.

(c) Co-localization of tdTomato with astrocytic marker GFAP. Sections were stained with antibodies against NeuN and GFAP. Arrowheads denote colocalization between tdTomato and GFAP. Scale bar: top, 200 μm; bottom, 30 μm.

(d) Representative Golgi staining images of primary and second/tertiary (sec/ter) dendrites in CA 1 pyramidal neurons of control and GFAP-CreER;Lrp4–/– mice. Scale bar, 10 μm.

(e) Quantitative analysis of data in (d). For primary dendrites, n = 47 segments, 47 neurons from 3 control mice; n = 40 segments, 40 neurons from 3 GFAP-CreER;Lrp4–/– mice, student’s t test, t(85) = 1.011, p = 0.3151. For sec/ter dendrites, n = 74 segments, 47 neurons from 3 control mice; n = 77 segments, 40 neurons from 3 GFAP-CreER;Lrp4–/– mice, student’s t test, t(149) = 1.434, p = 0.1537.

Data in e were presented as median with interquartile range, whiskers are the minimum and maximum; p > 0.05.

Supplementary Figure 6 Similar sEPSCs and PPR between control and Camk2 a-Lrp4–/– hippocampal pyramidal neurons.

(a) Cre activity was expressed in neurons in Camk2α::Cre mice. Shown were a hippocampal slice of Camk2α::Cre;tdTomato mice stained with neuronal marker NeuN (green) and astrocytic marker GFAP (blue) on the left. Images in the dotted areas (marked as a’ in CA1 and a’’ in CA3) were enlarged and shown on the right. Scale bar: left: 200 μm; right: 30 μm.

(b) Lrp4 level in the hippocampus was comparable between one-month-old control and Camk2α-Lrp4–/– mice. Shown were representative blots of three independent experiments with similar results. Full-length blots/gels are presented in Supplementary Figure 12.

(c) Quantitative analysis of data in (b). n = 6 hippocampi from 3 mice for each genotype. Lrp4 band density was normalized by the loading control β-actin; values of control mice were taken as 100%. Paired student’s t test, t(5) = 1.297, p = 0.2512.

(d), Diagrams showing recording of pyramidal neuron in whole-cell patch configuration. Note that blue color denotes Lrp4 mutant and red color denotes control neurons.

(e) Representative sEPCS traces of CA1 pyramidal neurons of control and Camk2α-Lrp4–/– mice. Scale bar: 2 s, 10 pA.

(f, g) No difference in sEPSC frequency (f) and amplitude (g). n = 10 neurons of 3 mice for both genotypes. For (f), student’s t test, t(18) = 0.6552, p = 0.5206; for (g), student’s t test, t(18) = 0.3296, p = 0.7455.

(h) No difference in paired-pulse ratios between control and mutant neurons. n = 6 neurons, 3 control mice; n = 8 neurons, 3 Camk2α-Lrp4–/– mice. Two-way ANOVA, F(1,35) = 0.02, p = 0.8843.

Data in (c and h) were presented as mean ± s.e.m; data in (f and g) were presented as median with interquartile range, whiskers are the minimum and maximum; p > 0.05.

Supplementary Figure 7 Characterization of processes and numbers of GFAP-Lrp4–/– astrocytes in vivo and in vitro.

(a) Cortical sections of control (top) and GFAP-Lrp4–/– (bottom) mice were stained for GFAP. Enlarged images of dotted area were shown on the right. SR: Stratum radiatum; SLM: stratum lacunosum- moleculare; ML: molecular layer. Scale bar: left, 150 μm; right, 30 μm.

(b) Representative traced astrocytes were superimposed with concentric circles in Sholl analysis. The radius interval between circles was 5 µm per step, ranging from 10 µm to 50 µm from the center of the astrocytic soma.

(c) Quantification of GFAP+ cells in SC region and Molecular Layers (ML). n = 6 slices of 3 control mice; n = 7 slices of 3 GFAP-Lrp4–/– mice. Two-way ANOVA, F(1,22) = 2.236, p = 0.149.

(d) Quantitative analysis of total process length of astrocytes. n = 32 astrocytes, 3 control mice; n = 31 astrocytes, 3 GFAP-Lrp4–/– mice. Student’s t test, t(61) = 1.339, p = 0.1856.

(e) Quantitative analysis of soma size of astrocytes. n = 70 astrocytes, 3 control mice; n = 65 astrocytes, 3 GFAP-Lrp4–/– mice. Student’s t test, t(133) = 0.6318, p = 0.5286.

(f) Quantitative analysis of process intersections. n = 32 astrocytes, 3 control mice; n = 31 astrocytes, 3 GFAP-Lrp4–/– mice. Two-way ANOVA, F(1,486) = 2.814, p = 0.0941.

(g) Representative images of neurons. Neurons were stained with MAP2 and DAPI, as indicated in green and blue, respectively. Scale bar: 50 μm.

(h) Quantitative analysis of MAP2+ cells. n = 8 coverslips for both genotypes. Student’s t test, t(14) = 0.2747, p = 0.7876.

(i) Representative images of astrocytes. Astrocytes were stained with GFAP and DAPI, as indicated in green and blue, respectively. Scale bar: 20 μm.

(j) Quantitative analysis of GFAP+ cells. n = 8 coverslip for control group; n = 9 coverslip for GFAP-Lrp4–/– group. Student’s t test, t(15) = 0.074, p = 0.9419.

Data in (c-e, h and j) were presented as median with interquartile range, whiskers are the minimum and maximum; data in (f) were presented as mean ± s.e.m; p > 0.05.

Supplementary Figure 8 Effects of pharmacological chemicals on sEPSC frequency.

Hippocampal slices of control mice were treated with different concentrations of indicated chemicals. sEPSCs were recorded in the CA1 pyramidal neurons. sEPSC frequency prior to treatment of each chemical was taken as 100% (Baseline). n = 6 and 8 for 10 µM and 100 µM, respectively for DL-AP5 (a), paired student’s t test, baseline vs 10 µM, t(5) = 2.311, p = 0.0688; baseline vs 100 µM, t(7) = 2.87, p = 0.024; n = 6 for both 1 µM and 10 µM, respectively for ATP (b), paired student’s t test, baseline vs 1 µM, t(6) = 2.26, p = 0.0645; baseline vs 10 µM, t(6) = 7.926, p = 0.0002; n = 6 and 8 for 10 µM and 300 µM, respectively for AOPCP (c), paired student’s t test, baseline vs 10 µM, t(5) = 1.943, p = 0.1097; baseline vs 300 µM, t(7) = 3.115, p = 0.017; n = 6 and 10 for 1 µM and 10 µM, respectively for Suramin (d), paired student’s t test, baseline vs 1 µM, t(5) = 1.604, p = 0.1697; baseline vs 10 µM, t(8) = 3.072, p = 0.0153; n = 5 and 7 for 10 nM and 800 nM, respectively for DPCPX (e), paired student’s t test, baseline vs 10 nM, t(4) = 1.012, p = 0.3688; baseline vs 800 nM, t(6) = 2.927, p = 0.0264; n = 6 and 8 for 1.5 µM and 5 µM, respectively for SCH58261 (f), paired student’s t test, baseline vs 1.5 µM, t(5) = 1.613, p = 0.1677; baseline vs 5 µM, t(7) = 2.556, p = 0.0378.

Data were presented as mean ± s.e.m; *, p < 0.05; **, p < 0.01.

Supplementary Figure 9 No effect of agrin on astrocyte proliferation.

(a) Lrp4 surface expression in astrocytes. Cultured astrocytes were treated with biotin. Biotin-labeled proteins were precipitated with Avidin agarose beads. Cell surface fraction and 10% cell lysates were probed for Lrp4 and TFR (transferin receptor protein). Shown were representative blots of two independent experiments with similar results. Full-length blots/gels are presented in Supplementary Figure 12.

(b) Agrin activation of MuSK in muscle cells. C2C12 myotubes were treated without or with agrin (100 ng/ml) for 20 min. MuSK was isolated by immunoprecipitation and probed with p-tyrosine antibody to detect phosphorylated MuSK. Shown were representative blots of three independent experiments with similar results. Full-length blots/gels are presented in Supplementary Figure 12.

(c) Quantitative analysis of data in (b). n = 3 times for control group, n = 4 times for agrin group. p-Tyrosine band density was normalized by total MuSK; values of control mice were taken as 100%. Student’s t test, t(5) = 3.137, p = 0.0257.

(d) C2C12 myotubes were stimulated with or without agrin (100 ng/ml) for 16 h. AChR clusters were stained by R-BTX. Arrows, AChR clusters. Scale bar: 50 μm.

(e) Quantitative analysis of data in (d). n = 8 images for each group. Student’s t test, t(14) = 6.966, p < 0.0001.

(f) Effect of agrin on proliferating astrocytes. Representative images of control (left) and GFAP-Lrp4–/– astrocytes (right) with or without agrin treatment. Astrocytes were stained with Ki67 and GFAP, respectively. Scale bar: 40 μm.

(g) Quantitative analysis of Ki67+ cells. n = 12 coverslip for each group. Data were presented as mean ± s.e.m. One-way ANOVA, F(1,44) = 0.188, p = 0.904.

Data were presented as mean ± s.e.m; *, p < 0.05; **, p < 0.01.

Supplementary Figure 10 Characterization of GFAP-Lrp4–/– mice in elevated plus maze, light/dark box, and water maze.

(a) Representative traces of elevated plus maze tests. Scale bar: 5 cm.

(b, c, d) No difference in duration in the open arms, number of entries to open arms, latency to open arms. n = 9 mice for each genotype. For (b), student’s t test, t(16) = 0.1493, p = 0.8832; for (c), student’s t test, t(16) = 0.2916, p = 0.7744; for (d), student’s t test, t(16) = 0.8321, p = 0.4176.

(e) Representative traces in the light/dark box test. Scale bar: 5 cm.

(f, g, h) No difference in duration in light box, number of entries to light box, latency to light box. n = 11 mice for each genotype. For (f), student’s t test, t(20) = 0.8256, p = 0.4188; for (g), student’s t test, t(20) = 0.701, p = 0.4914; for (h), student’s t test, t(20) = 0.6747, p = 0.5076.

(i) Quantitative analysis of velocity in Morris water maze of two genotypes. n = 9 mice for each genotype. Student’s t test; t(16) = 0.6239, p = 0.5415.

(j) Time-course analysis of duration in the N30 area in probe test. n = 9 mice for each genotype. Repeated two-way ANOVA, for genotype and time interaction: F(5, 80) = 1.032, p = 0.4047.

(k) Time-course analysis of platform crossing numbers in probe test. n = 9 mice for each genotype. Repeated two-way ANOVA, for genotype and time interaction: F(5, 80) = 0.1515, p = 0.979.

Data were presented as mean ± s.e.m; p > 0.05.

Supplementary Figure 11 A working model for Lrp4 in astrocytes to modulate glutamatergic transmission.

Lrp4 in astrocytes serves as a receptor for agrin and controls ATP release from astrocytes. ATP is hydrolyzed to adenosine, which suppresses glutamate release by activating A1 purinergic receptors.

Supplementary Figure 12 Full-length pictures of the blots presented in the main figures and supplementary figures.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–12 (PDF 5340 kb)

Supplementary Methods Checklist

(PDF 605 kb)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sun, XD., Li, L., Liu, F. et al. Lrp4 in astrocytes modulates glutamatergic transmission. Nat Neurosci 19, 1010–1018 (2016). https://doi.org/10.1038/nn.4326

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nn.4326

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing