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:

Tau induces PSD95–neuronal NOS uncoupling and neurovascular dysfunction independent of neurodegeneration

Abstract

Cerebrovascular abnormalities have emerged as a preclinical manifestation of Alzheimer’s disease and frontotemporal dementia, diseases characterized by the accumulation of hyperphosphorylated forms of the microtubule-associated protein tau. However, it is unclear whether tau contributes to these neurovascular alterations independent of neurodegeneration. We report that mice expressing mutated tau exhibit a selective suppression of neural activity-induced cerebral blood flow increases that precedes tau pathology and cognitive impairment. This dysfunction is attributable to a reduced vasodilatation of intracerebral arterioles and is reversible by reducing tau production. Mechanistically, the failure of neurovascular coupling involves a tau-induced dissociation of neuronal nitric oxide synthase (nNOS) from postsynaptic density 95 (PSD95) and a reduced production of the potent vasodilator nitric oxide during glutamatergic synaptic activity. These data identify glutamatergic signaling dysfunction and nitric oxide deficiency as yet-undescribed early manifestations of tau pathobiology, independent of neurodegeneration, and provide a mechanism for the neurovascular alterations observed in the preclinical stages of tauopathies.

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

Fig. 1: Neurovascular coupling is selectively disrupted in tau mice before tau pathology and cognitive impairment.
Fig. 2: Suppression of neurovascular coupling in tau mice results from impaired vasodilatation of intracerebral arterioles.
Fig. 3: Spontaneous and evoked neural activity and NMDA-induced Ca2+ increase are not altered in PS19 mice.
Fig. 4: Suppressing tau production rescues neurovascular coupling and prevents cortical atrophy and cognitive deficits in older rTg4510 mice.
Fig. 5: The NMDA-dependent component of functional hyperemia is selectively suppressed in tau mice.
Fig. 6: The NO-dependent component of functional hyperemia is attenuated in tau mice.
Fig. 7: Suppression of NMDAR-induced neuronal NO production and dissociation of nNOS from PSD95 in tau mice.
Fig. 8: Tau disrupts the PSD95–nNOS association by binding to PSD95.

Similar content being viewed by others

Data availability

All the data supporting the conclusions of the current study are presented in the figures. If necessary, the data that support the findings of this study are available from the corresponding authors upon reasonable request. There are no restrictions on data availability. Source data are provided with this paper.

Code availability

No code was used for the study.

References

  1. Cummings, J., Lee, G., Ritter, A. & Zhong, K. Alzheimer’s disease drug development pipeline: 2018. Alzheimers Dement. (NY) 4, 195–214 (2018).

    Article  Google Scholar 

  2. Henstridge, C. M., Hyman, B. T. & Spires-Jones, T. L. Beyond the neuron–cellular interactions early in Alzheimer disease pathogenesis. Nat. Rev. Neurosci. 20, 94–108 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Knopman, D. S. Lowering of amyloid-beta by β-secretase inhibitors—some informative failures. N. Engl. J. Med. 380, 1476–1478 (2019).

    Article  CAS  PubMed  Google Scholar 

  4. Boyle, P. A. et al. Attributable risk of Alzheimer’s dementia attributed to age-related neuropathologies. Ann. Neurol. 85, 114–124 (2019).

    Article  CAS  PubMed  Google Scholar 

  5. Iadecola, C. & Gottesman, R. F. Cerebrovascular alterations in Alzheimer disease. Circ. Res. 123, 406–408 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Iturria-Medina, Y. et al. Early role of vascular dysregulation on late-onset Alzheimer’s disease based on multifactorial data-driven analysis. Nat. Commun. 7, 11934 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Rabin, J. S. et al. Vascular risk and beta-amyloid are synergistically associated with cortical tau. Ann. Neurol. 85, 272–279 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Iadecola, C. The neurovascular unit coming of age: a journey through neurovascular coupling in health and disease. Neuron 96, 17–42 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Scheltens, P. et al. Alzheimer’s disease. Lancet 388, 505–517 (2016).

    Article  CAS  PubMed  Google Scholar 

  10. Dopper, E. G. et al. Cerebral blood flow in presymptomatic MAPT and GRN mutation carriers: a longitudinal arterial spin labeling study. Neuroimage Clin. 12, 460–465 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  11. Kurata, T. et al. PSP as distinguished from CBD, MSA-P and PD by clinical and imaging differences at an early stage. Intern. Med. 50, 2775–2781 (2011).

    Article  CAS  PubMed  Google Scholar 

  12. Lunau, L. et al. Presymptomatic cerebral blood flow changes in CHMP2B mutation carriers of familial frontotemporal dementia (FTD-3), measured with MRI. BMJ Open 2, e000368 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  13. Iadecola, C. The pathobiology of vascular dementia. Neuron 80, 844–866 (2013).

    Article  CAS  PubMed  Google Scholar 

  14. Sweeney, M. D., Kisler, K., Montagne, A., Toga, A. W. & Zlokovic, B. V. The role of brain vasculature in neurodegenerative disorders. Nat. Neurosci. 21, 1318–1331 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Brenman, J. E. et al. Interaction of nitric oxide synthase with the postsynaptic density protein PSD-95 and α1-syntrophin mediated by PDZ domains. Cell 84, 757–767 (1996).

    Article  CAS  PubMed  Google Scholar 

  16. Christopherson, K. S., Hillier, B. J., Lim, W. A. & Bredt, D. S. PSD-95 assembles a ternary complex with the N-methyl-d-aspartic acid receptor and a bivalent neuronal NO synthase PDZ domain. J. Biol. Chem. 274, 27467–27473 (1999).

    Article  CAS  PubMed  Google Scholar 

  17. Kornau, H. C., Schenker, L. T., Kennedy, M. B. & Seeburg, P. H. Domain interaction between NMDA receptor subunits and the postsynaptic density protein PSD-95. Science 269, 1737–1740 (1995).

    Article  CAS  PubMed  Google Scholar 

  18. Girouard, H. et al. NMDA receptor activation increases free radical production through nitric oxide and NOX2. J. Neurosci. 29, 2545–2552 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Mishra, A. et al. Astrocytes mediate neurovascular signaling to capillary pericytes but not to arterioles. Nat. Neurosci. 19, 1619–1627 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Park, L. et al. Key role of tissue plasminogen activator in neurovascular coupling. Proc. Natl Acad. Sci. USA 105, 1073–1078 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Bennett, R. E. et al. Tau induces blood vessel abnormalities and angiogenesis-related gene expression in P301L transgenic mice and human Alzheimer’s disease. Proc. Natl Acad. Sci. USA 115, E1289–E1298 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Blair, L. J. et al. Tau depletion prevents progressive blood–brain barrier damage in a mouse model of tauopathy. Acta Neuropathol. Commun. 3, 8 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Ittner, A. & Ittner, L. M. Dendritic tau in Alzheimer’s disease. Neuron 99, 13–27 (2018).

    Article  CAS  PubMed  Google Scholar 

  24. Mondragon-Rodriguez, S. et al. Interaction of endogenous tau protein with synaptic proteins is regulated by N-methyl-d-aspartate receptor-dependent tau phosphorylation. J. Biol. Chem. 287, 32040–32053 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Yoshiyama, Y. et al. Synapse loss and microglial activation precede tangles in a P301S tauopathy mouse model. Neuron 53, 337–351 (2007).

    Article  CAS  PubMed  Google Scholar 

  26. Santacruz, K. et al. Tau suppression in a neurodegenerative mouse model improves memory function. Science 309, 476–481 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Hoover, B. R. et al. Tau mislocalization to dendritic spines mediates synaptic dysfunction independently of neurodegeneration. Neuron 68, 1067–1081 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Ramsden, M. et al. Age-dependent neurofibrillary tangle formation, neuron loss, and memory impairment in a mouse model of human tauopathy (P301L). J. Neurosci. 25, 10637–10647 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Niwa, K., Haensel, C., Ross, M. E. & Iadecola, C. Cyclooxygenase-1 participates in selected vasodilator responses of the cerebral circulation. Circ. Res. 88, 600–608 (2001).

    Article  CAS  PubMed  Google Scholar 

  30. Iadecola, C. Does nitric oxide mediate the increases in cerebral blood flow elicited by hypercapnia? Proc. Natl Acad. Sci. USA 89, 3913–3916 (1992).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Iadecola, C., Pelligrino, D. A., Moskowitz, M. A. & Lassen, N. A. Nitric oxide synthase inhibition and cerebrovascular regulation. J. Cereb. Blood Flow. Metab. 14, 175–192 (1994).

    Article  CAS  PubMed  Google Scholar 

  32. Uekawa, K. et al. Obligatory role of EP1 receptors in the increase in cerebral blood flow produced by hypercapnia in the mice. PLoS ONE 11, e0163329 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Fa, M. et al. Extracellular tau oligomers produce an immediate impairment of LTP and memory. Sci. Rep. 6, 19393 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Yamada, K. et al. In vivo microdialysis reveals age-dependent decrease of brain interstitial fluid tau levels in P301S human tau transgenic mice. J. Neurosci. 31, 13110–13117 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Brochner, C. B., Holst, C. B. & Mollgard, K. Outer brain barriers in rat and human development. Front. Neurosci. 9, 75 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  36. Sykova, E. Diffusion properties of the brain in health and disease. Neurochem. Int. 45, 453–466 (2004).

    Article  CAS  PubMed  Google Scholar 

  37. Chen, B. R., Kozberg, M. G., Bouchard, M. B., Shaik, M. A. & Hillman, E. M. A critical role for the vascular endothelium in functional neurovascular coupling in the brain. J. Am. Heart Assoc. 3, e000787 (2014).

    PubMed  PubMed Central  Google Scholar 

  38. Longden, T. A. et al. Capillary K+-sensing initiates retrograde hyperpolarization to increase local cerebral blood flow. Nat. Neurosci. 20, 717–726 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Lecrux, C. et al. Pyramidal neurons are “neurogenic hubs” in the neurovascular coupling response to whisker stimulation. J. Neurosci. 31, 9836–9847 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Buerk, D. G., Ances, B. M., Greenberg, J. H. & Detre, J. A. Temporal dynamics of brain tissue nitric oxide during functional forepaw stimulation in rats. Neuroimage 18, 1–9 (2003).

    Article  PubMed  Google Scholar 

  41. Koizumi, K. et al. Apoε4 disrupts neurovascular regulation and undermines white matter integrity and cognitive function. Nat. Commun. 9, 3816 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Sattler, R. et al. Specific coupling of NMDA receptor activation to nitric oxide neurotoxicity by PSD-95. Protein Sci. 284, 1845–1848 (1999).

    CAS  Google Scholar 

  43. Kopeikina, K. J. et al. Synaptic alterations in the rTg4510 mouse model of tauopathy. J. Comp. Neurol. 521, 1334–1353 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Warmus, B. A. et al. Tau-mediated NMDA receptor impairment underlies dysfunction of a selectively vulnerable network in a mouse model of frontotemporal dementia. J. Neurosci. 34, 16482–16495 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  45. Gamache, J. et al. Factors other than hTau overexpression that contribute to tauopathy-like phenotype in rTg4510 mice. Nat. Commun. 10, 2479 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  46. Goodwin, L. O. et al. Large-scale discovery of mouse transgenic integration sites reveals frequent structural variation and insertional mutagenesis. Genome Res. 29, 494–505 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Hardingham, N., Dachtler, J. & Fox, K. The role of nitric oxide in pre-synaptic plasticity and homeostasis. Front. Cell Neurosci. 7, 190 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Garthwaite, J. NO as a multimodal transmitter in the brain: discovery and current status. Br. J. Pharmacol. 176, 197–211 (2019).

    Article  CAS  PubMed  Google Scholar 

  49. Zhu, J., Shang, Y. & Zhang, M. Mechanistic basis of MAGUK-organized complexes in synaptic development and signalling. Nat. Rev. Neurosci. 17, 209–223 (2016).

    Article  CAS  PubMed  Google Scholar 

  50. Karp, N. A. et al. Applying the ARRIVE guidelines to an in vivo database. PLoS Biol. 13, e1002151 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  51. Franklin, K. B. J. & Paxinos, G. The Mouse Brain in Stereotaxic Coordinates (Academic Press, 1997).

  52. Jackman, K. et al. Progranulin deficiency promotes post-ischemic blood–brain barrier disruption. J. Neurosci. 33, 19579–19589 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Park, L. et al. Age-dependent neurovascular dysfunction and damage in a mouse model of cerebral amyloid angiopathy. Stroke 45, 1815–1821 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Park, L. et al. The key role of transient receptor potential melastatin-2 channels in amyloid-β-induced neurovascular dysfunction. Nat. Commun. 5, 5318 (2014).

    Article  CAS  PubMed  Google Scholar 

  55. Iadecola, C. Nitric oxide participates in the cerebrovasodilation elicited from cerebellar fastigial nucleus. Am. J. Physiol. 263, R1156–R1161 (1992).

    CAS  PubMed  Google Scholar 

  56. Park, L. et al. Scavenger receptor CD36 is essential for the cerebrovascular oxidative stress and neurovascular dysfunction induced by amyloid-beta. Proc. Natl Acad. Sci. USA 108, 5063–5068 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Cruz Hernandez, J. C. et al. Neutrophil adhesion in brain capillaries reduces cortical blood flow and impairs memory function in Alzheimer’s disease mouse models. Nat. Neurosci. 22, 413–420 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Shih, A. Y., Mateo, C., Drew, P. J., Tsai, P. S. & Kleinfeld, D. A polished and reinforced thinned-skull window for long-term imaging of the mouse brain. J. Vis. Exp. 7, 3742 (2012).

    Google Scholar 

  59. Dunn, A. K., Bolay, H., Moskowitz, M. A. & Boas, D. A. Dynamic imaging of cerebral blood flow using laser speckle. J. Cereb. Blood Flow. Metab. 21, 195–201 (2001).

    Article  CAS  PubMed  Google Scholar 

  60. Park, L. et al. Exogenous NADPH increases cerebral blood flow through NADPH oxidase-dependent and -independent mechanisms. Arterioscler. Thromb. Vasc. Biol. 24, 1860–1865 (2004).

    Article  CAS  PubMed  Google Scholar 

  61. Kazama, K., Wang, G., Frys, K., Anrather, J. & Iadecola, C. Angiotensin II attenuates functional hyperemia in the mouse somatosensory cortex. Am. J. Physiol. Heart Circ. Physiol. 285, H1890–H1899 (2003).

    Article  CAS  PubMed  Google Scholar 

  62. Coleman, C. G. et al. Chronic intermittent hypoxia induces NMDA receptor-dependent plasticity and suppresses nitric oxide signaling in the mouse hypothalamic paraventricular nucleus. J. Neurosci. 30, 12103–12112 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Wang, G. et al. Angiotensin II slow-pressor hypertension enhances NMDA currents and NOX2-dependent superoxide production in hypothalamic paraventricular neurons. Am. J. Physiol. Regul. Integr. Comp. Physiol. 304, R1096–R1106 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Kawano, T. et al. Prostaglandin E2 EP1 receptors: downstream effectors of COX-2 neurotoxicity. Nat. Med. 12, 225–229 (2006).

    Article  CAS  PubMed  Google Scholar 

  65. Machida, S. et al. Cycloamylose as an efficient artificial chaperone for protein refolding. FEBS Lett. 486, 131–135 (2000).

    Article  CAS  PubMed  Google Scholar 

  66. Ittner, L. M. et al. Dendritic function of tau mediates amyloid-β toxicity in Alzheimer’s disease mouse models. Cell 142, 387–397 (2010).

    Article  CAS  PubMed  Google Scholar 

  67. Peng, H. M., Morishima, Y., Pratt, W. B. & Osawa, Y. Modulation of heme/substrate binding cleft of neuronal nitric-oxide synthase (nNOS) regulates binding of Hsp90 and Hsp70 proteins and nNOS ubiquitination. J. Biol. Chem. 287, 1556–1565 (2012).

    Article  CAS  PubMed  Google Scholar 

  68. Hochrainer, K. et al. The ubiquitin ligase HERC3 attenuates NF-κB-dependent transcription independently of its enzymatic activity by delivering the RelA subunit for degradation. Nucleic Acids Res. 43, 9889–9904 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Faraco, G. et al. Perivascular macrophages mediate the neurovascular and cognitive dysfunction associated with hypertension. J. Clin. Invest. 126, 4674–4689 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

This work was supported by NIH grants R01-NS37853 (to C.I.), R01-NS097805 (to L.P.) and R01-NS109588 (to K.H.), the Japan Heart Foundation/Bayer Research Grant Abroad (to Y.H.), The Uehara Memorial Foundation Research Fellowship (to Y.H.), the Japan Society for the Promotion of Science Overseas Research Fellowship (to Y.H.), and American Heart Association Postdoctoral Fellowship 20POST35120063 (to S.J.A.). We thank C. B. Schaffer for helpful suggestions and editing. Support from the Feil Family Foundation is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Contributions

L.P., K.H. and Y.H. conducted the experiments and performed the data analyses. S.J.A. and A.A. conducted the 2PM and LSI experiments. G.W. performed the NO production and Ca2+ imaging experiments. J.S. and K.U. contributed to the histology experiments. I.B. and V.P. contributed to the immunoprecipitation and western blotting experiments. D.E. and D.A. provided WT rTau. L.P., K.H., P.Z., J.A. and C.I. supervised the research. L.P., K.H. and C.I. provided funding. L.P. and C.I. wrote the manuscript.

Corresponding authors

Correspondence to Laibaik Park or Costantino Iadecola.

Ethics declarations

Competing interests

C.I. serves on the advisory board of Broadview Ventures. The other authors have no conflicts to declare.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data

Extended Data Fig. 1 Resting CBF and thickness in PS19, rTg4510, and WT mice at 2-3 months of age.

Resting CBF and thickness were bilaterally assessed by ASL- and T2-MRI, respectively, in neocortex, entorhinal cortex, and hippocampus at the bregma level from -1.22 to -1.70 mm. a, b, A small reduction in entorhinal (A) and hippocampal (B) CBF is found in PS19 mice and not in rTg4510 mice compared to age-matched WT mice, while entorhinal cortex thickness (A) and hippocampal volume (B) are comparable in both strains. N = 10/group; one-way analysis of variance (ANOVA) with Tukey’s test for multiple comparisons. c, d, Resting CBF (C) and thickness (D) in the neocortex, entorhinal cortex, and hippocampus are comparable in rTg4510 mice, expressing both tau (Tau+) and transactivator (TA+), and their littermates (WT; Tau+ only; TA+ only). Scale bar=1 mm; N = 5/group; one-way ANOVA with Tukey’s test. Data are presented as mean ± SEM. See Source Data 7 for statistical parameters.

Source data

Extended Data Fig. 2 No neurofibrillary tangles are observed in 2-3-month-old PS19, rTg4510, and WT mice.

Neurofibrillary tangles, assessed by the thioflavin-S stain, are not observed in the somatosensory cortex of 2-3-month-old WT (a–c), PS19 (d), and rTg4510 (g) mice, but phosphorylated tau (AT-8) is observed in PS19 (e, f) and rTg4510 (h, i). Images are representative of 3 independent experiments, each including 5 mice/group.

Extended Data Fig. 3 Microvessels, microglia/macrophages, and astrocytes in the neocortex of 2-3-month-old tau mice.

Neocortical cluster of differentiation (CD) 31 (CD31+) microvessels (a; N = 5/group) and Iba1+ microglia/macrophages (b; N = 5/group) are comparable in PS19, rTg4510, and WT mice at the age of 2-3 months, but enhanced astrogliosis (GFAP+ cells) is found in rTg4510 mice, compared to PS19 and WT mice (c; N = 5/group). One-way ANOVA and Tukey’s test. Data are presented as mean ± SEM. See Source Data 8 for statistical parameters.

Source data

Extended Data Fig. 4 No neuronal loss occurs in PS19 and rTg4510 mice, but mislocalized phosphorylated tau is observed in somatodendritic compartments.

The number of neurons (NeuN+) is comparable in PS19, rTg4510, and WT mice (a, see quantification in Fig. 1c). As anticipated, in PS19 and rTg4510 mice phosphorylated tau (AT-8+) is co-localized with neurons (NeuN+) (A) and axons (myelin basic protein, MBP+) (b). Phosphorylated tau is also observed in dendritic spines (MAP2+) (c), indicating displacement of tau to somatodendritic compartments. Images are representative of 3 independent experiments, each including 5 mice/group.

Extended Data Fig. 5 Locomotor activity and neurovascular response in 2-3-month-PS19 and rTg4510 mice.

No changes in locomotor activity are observed in novel object recognition and Y-maze tests of rTg4510, compared to WT mice. N = 5 for novel object; N = 6 for Y-maze. b, The increases in CBF-LDF induced in the whisker barrel cortex by mechanical stimulation of the facial whiskers were markedly attenuated in PS19 mice also under the isoflurane anesthesia regimen used in ASL-MRI studies. N = 5/group; two-tailed unpaired t-test. c, The increases in CBF-LDF produced by the endothelium-dependent (bradykinin or A23187) and -independent (SNAP, adenosine, or hypercapnia) vasodilators are comparable in PS19, rTg4510, and WT mice. N = 5/group. d, Recombinant full-length mutant (2N4R, P301L; 5 µM) or WT (2N4R; 5 µM) tau has no effect on CBF response induced by acetylcholine or adenosine. N = 5/group. Data are presented as mean ± SEM. One-way ANOVA and Tukey’s test.

Extended Data Fig. 6 Suppressing tau production with doxycycline reduces P301L tau expression and prevents neuronal loss, tau accumulation in 7-8-month-old rTg4510 mice.

a, Doxycycline treatment (tau off) for 3-4 months reduces neuronal loss in rTg4510 mice, compared to rTg4510 mice fed control diet (con or tau on). Treatment with doxycycline (doxy) has no effect in WT mice (See Fig. 4c for quantification). b, Suppressing tau production reduces AT-8+ tau levels, but not thioflavin-S+ neurofibrillary tangles. Arrows indicate co-localization between thioflavin-S+ neurofibrillary tangles and AT8+ tau, and asterisks denote strong thioflavin-S+ neurofibrillary tangles with faint or no AT8+ tau. Representative pictures from N = 5 mice/group. c, Doxycycline treatment reduces human tau P301L mRNA, but has no effect on mouse tau mRNA. N = 5 for WT in mouse tau mRNA and for rTg4510 (tau on) in human tau mRNA; N = 4 for rTg4510 in mouse tau mRNA; N = 6 for rTg4510 (tau off) in human tau mRNA. Images are representative of 3 independent experiments, each including 5 mice/group. Two-tailed unpaired t-test. Data are presented as mean ± SEM.

Extended Data Fig. 7 Suppressing tau production with doxycycline prevents cognitive deficits in 7-8-month-old rTg4510 mice.

a, Suppressing tau production with doxycycline (tau off or doxy), compared with control diet (tau on or con), prevents the CBF reduction, assessed by ASL-MRI. N = 7/group; two-way ANOVA with Tukey’s test. b, c, Suppressing tau production prevents cognitive deficits, assessed by novel object recognition (b; N = 10 for WT tau on & off and rTg4510 tau on; N = 9 for rTg4510 tau off) and Y-maze test (c; N = 10/group), but has no effect on locomotor activity, as reflected by distance traveled (b) or number of arm entries (c). Locomotor activity of rTg4510 mice in the novel object test seems more variable (b). Data are presented as mean ± SEM. Two-way ANOVA with Tukey’s test. d, The CBF increase induced by neocortical superfusion of the NO donor SNAP or adenosine is not altered in 7-8 month-old rTg4510 mice. N = 5/group; two-way ANOVA with Tukey’s test. Data are presented as mean ± SEM. See Source Data 9 for statistical parameters.

Source data

Extended Data Fig. 8 Aquaporin-4 immunoreactivity and astrogliosis are unaffected by suppressing tau production with doxycycline in 7-8-month-old rTg4510 mice.

a, AQP-4 immunoreactivity, which labels astrocytic end-feet, was not disrupted in rTg4510 mice with or without doxycycline. N = 5/group; two-way ANOVA with Tukey’s test; p = 0.2420 between WT con & WT doxy, p = 0.9985 between WT con & rTg4510 tau on, p = 0.8585 between WT con & rTg4510 tau off, p = 0.1881 between WT doxy & rTg4510 tau on, p = 0.6502 between WT doxy & rTg4510 tau off, and p = 0.7804 between rTg4510 tau on & off. Data are presented as mean ± SEM. b, The astrogliosis (GFAP+ cells) observed in rTg4510 mice was not reduced with tau suppression. Data are presented as mean ± SEM. N = 5/group; two-way ANOVA with Tukey’s test. See Source Data 10 for statistical parameters.

Source data

Extended Data Fig. 9 MK-801 and TTX effect on CBF, L-NNA effect on NMDA-induced NO production and expression of NMDA receptor subunits in 2-3-month-old rTg4510 mice.

a, MK-801 and/or TTX have no effect on resting CBF or CBF response produced by neocortical superfusion of acetylcholine or adenosine. N = 5/group. Data are presented as mean ± SEM. b, NMDAR subunits mRNA levels are comparable in WT and rTg4510 mice. N = 5/group. Data are presented as mean ± SEM. c, Treatment with the NOS inhibitor L-NNA prevents NMDA-induced NO production in isolated cortical neurons from WT mice. Data are presented as mean ± SEM. N = 6/group; one-way ANOVA with Tukey’s test. See Source Data 11 for statistical parameters.

Source data

Extended Data Fig. 10 NMDAR-related proteins in PS19 and rTg4510 mice, effects of WT tau on nNOS-PSD95 coupling, and putative mechanisms of the effect of tau on neurovascular function.

a–d, Levels and kinase activity of NMDAR-related proteins are not altered in 2-3 month-old PS19 and rTg4510 mice. a, Protein levels of nNOS, GluN2B and PSD95 are unaltered in synaptosomal preparations of PS19 (PS) compared to WT mice. N = 3/group. Data are presented as mean ± SEM. b, Protein levels of nNOS, GluN2B and PSD95 in PS19 (PS) and WT mice are comparable. The presynaptic marker synaptophysin (SYP) and MEK1/2 were used as membrane (MEM) and cytosolic (CYT) markers, respectively. N = 7/group. Data are presented as mean ± SEM. c, As in PS19 mice, nNOS, GluN2B and PSD95 protein levels are unchanged in PSD preparations of rTg4510 (Tg) mice compared to WT mice. N = 4/group. Data are presented as mean ± SEM. d, CaMKIIα level and activity quantified with reference to GAPDH, synaptophysin (SYP), or PSD95 associated with cytoplasm (CYT), membrane (MEM), and/or PSD are not altered in PS19 mice, compared to WT. N = 3/group. Data are presented as mean ± SEM. Immunoblots in a-d are cropped; full gel pictures are shown in Source Data 12. e-g, WT tau impairs binding of nNOS to PSD95 through association with PSD95. e, WT tau over-expressed in HEK293T cells is susceptible to phosphatase treatment, and thus hyperphosphorylated. N = 3/group; unpaired t-test. f, WT tau co-expression disrupts binding of nNOS to precipitated PSD95. N = 8/group; two-tailed unpaired t-test. g, Exogenously expressed WT tau interacts with PSD95 in HEK293T cells. A representative blot from N = 3/group is shown. Data are presented as mean ± SEM. Units for markers on the immunoblots in a-g are kDa. Immunoblots in e-g are cropped; full gel pictures are shown in Source Data 12. h, Putative mechanism by which tau induces a deficit in neuronal NO and neurovascular dysfunction: pathogenic tau (p-tau) binds to PSD95 and prevents its association with nNOS (1) and the resulting suppression in the NO production evoked by glutamatergic synaptic activity (2) dampens the NO dependent component of the increase in CBF produced by activation (3).

Supplementary information

Supplementary Information

Supplementary Tables 1 and 2: primary antibody list and PCR primers.

Reporting Summary

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 5

Statistical source data.

Source Data Fig. 6

Statistical source data.

Source Data Fig. 7d–f

Unprocessed/uncropped immunoblots.

Source Data Fig. 8

Unprocessed/uncropped immunoblots.

Source Data Extended Data Fig. 1

Statistical source data.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 8

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10a–g

Unprocessed/uncropped immunoblots.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Park, L., Hochrainer, K., Hattori, Y. et al. Tau induces PSD95–neuronal NOS uncoupling and neurovascular dysfunction independent of neurodegeneration. Nat Neurosci 23, 1079–1089 (2020). https://doi.org/10.1038/s41593-020-0686-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41593-020-0686-7

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