リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

大学・研究所にある論文を検索できる 「A light-induced small G-protein gem limits the circadian clock phase-shift magnitude by inhibiting voltage-dependent calcium channels」の論文概要。リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

コピーが完了しました

URLをコピーしました

論文の公開元へ論文の公開元へ
書き出し

A light-induced small G-protein gem limits the circadian clock phase-shift magnitude by inhibiting voltage-dependent calcium channels

Matsuo, Masahiro Seo, Kazuyuki Taruno, Akiyuki Mizoro, Yasutaka Yamaguchi, Yoshiaki Doi, Masao Nakao, Rhyuta Kori, Hiroshi Abe, Takaya Ohmori, Harunori Tominaga, Keiko Okamura, Hitoshi 京都大学 DOI:10.1016/j.celrep.2022.110844

2022.05

概要

Calcium signaling is pivotal to the circadian clockwork in the suprachiasmatic nucleus (SCN), particularly in rhythm entrainment to environmental light-dark cycles. Here, we show that a small G-protein Gem, an endogenous inhibitor of high-voltage-activated voltage-dependent calcium channels (VDCCs), is rapidly induced by light in SCN neurons via the calcium (Ca²⁺)-mediated CREB/CRE transcriptional pathway. Gem attenuates light-induced calcium signaling through its interaction with VDCCs. The phase-shift magnitude of locomotor activity rhythms by light, at night, increases in Gem-deficient (Gem⁻/⁻) mice. Similarly, in SCN slices from Gem⁻/⁻ mice, depolarizing stimuli induce larger phase shifts of clock gene transcription rhythms that are normalized by the application of an L-type VDCC blocker, nifedipine. Voltage-clamp recordings from SCN neurons reveal that Ca²⁺ currents through L-type channels increase in Gem⁻/⁻ mice. Our findings suggest that transcriptionally activated Gem feeds back to suppress excessive light-evoked L-type VDCC activation, adjusting the light-induced phase-shift magnitude to an appropriate level in mammals.

この論文で使われている画像

関連論文

参考文献

Allen, C.N., Nitabach, M.N., and Colwell, C.S. (2017). Membrane currents,

gene expression, and circadian clocks. Cold Spring Harb. Perspect. Biol. 9,

a027714. https://doi.org/10.1101/cshperspect.a027714.

Asai, M., Yamaguchi, S., Isejima, H., Jonouchi, M., Moriya, T., Shibata, S., Kobayashi, M., and Okamura, H. (2001). Visualization of mPer1 transcription

in vitro. Curr. Biol. 11, 1524–1527. https://doi.org/10.1016/s0960-9822(01)

00445-6.

Be´guin, P., Nagashima, K., Gonoi, T., Shibasaki, T., Takahashi, K., Kashima,

Y., Ozaki, N., Geering, K., Iwanaga, T., and Seino, S. (2001). Regulation of

Ca2+ channel expression at the cell surface by the small G-protein kir/Gem.

Nature 411, 701–706. https://doi.org/10.1038/35079621.

Be´guin, P., Ng, Y.J.A., Krause, C., Mahalakshmi, R.N., Ng, M.Y., and Hunziker,

W. (2007). RGK small GTP-binding proteins interact with the nucleotide kinase

domain of Ca2+-channel beta-subunits via an uncommon effector binding

domain. J. Biol. Chem. 282, 11509–11520. https://doi.org/10.1074/jbc.

M606423200.

Buraei, Z., Lumen, E., Kaur, S., and Yang, J. (2015). RGK regulation of voltagegated calcium channels. Sci. China Life Sci. 58, 28–38. https://doi.org/10.

1007/s11427-014-4788-x.

€ning, J.C., Sun, Z.,

Chang, L., Zhang, J., Tseng, Y.H., Xie, C.Q., Ilany, J., Bru

Zhu, X., Cui, T., Youker, K.A., et al. (2007). Rad GTPase deficiency leads to cardiac hypertrophy. Circulation 116, 2976–2983. https://doi.org/10.1161/circulationaha.107.707257.

Chen, H., Puhl, H.L., 3rd, Niu, S.L., Mitchell, D.C., and Ikeda, S.R. (2005).

Expression of Rem2, an RGK family small GTPase, reduces N-type calcium

current without affecting channel surface density. J. Neurosci. 25, 9762–

9772. https://doi.org/10.1523/jneurosci.3111-05.2005.

Chuhma, N., Koyano, K., and Ohmori, H. (2001). Synchronisation of neurotransmitter release during postnatal development in a calyceal presynaptic terminal of rat. J. Physiol. 530, 93–104. https://doi.org/10.1111/j.1469-7793.

2001.0093m.x.

Cohen, L., Mohr, R., Chen, Y.Y., Huang, M., Kato, R., Dorin, D., Tamanoi, F.,

Goga, A., Afar, D., Rosenberg, N., et al. (1994). Transcriptional activation of

a ras-like gene (kir) by oncogenic tyrosine kinases. Proc. Natl. Acad. Sci.

U S A 91, 12448–12452. https://doi.org/10.1073/pnas.91.26.12448.

Colecraft, H.M. (2020). Designer genetically encoded voltage-dependent calcium channel inhibitors inspired by RGK GTPases. J. Physiol. 598, 1683–1693.

https://doi.org/10.1113/jp276544.

Colwell, C.S. (2011). Linking neural activity and molecular oscillations in the

SCN. Nat. Rev. Neurosci. 12, 553–569. https://doi.org/10.1038/nrn3086.

Colwell, C.S., and Menaker, M. (1992). NMDA as well as non-NMDA receptor

antagonists can prevent the phase-shifting effects of light on the circadian system of the golden hamster. J. Biol. Rhythms 7, 125–136. https://doi.org/10.

1177/074873049200700204.

Colwell, C.S., Whitmore, D., Michel, S., and Block, G.D. (1994). Calcium plays

a central role in phase shifting the ocular circadian pacemaker of Aplysia.

J. Comp. Physiol. A 175, 415–423. https://doi.org/10.1007/bf00199249.

Crosio, C., Cermakian, N., Allis, C.D., and Sassone-Corsi, P. (2000). Light induces chromatin modification in cells of the mammalian circadian clock.

Nat. Neurosci. 3, 1241–1247. https://doi.org/10.1038/81767.

Daan, S., and Pittendrigh, C.S. (1976). A Functional analysis of circadian pacemakers in nocturnal rodents II. The variability of phase response curves.

J. Comp. Physiol. A 106, 253–266. https://doi.org/10.1007/bf01417857.

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

Report

Doi, M., Ishida, A., Miyake, A., Sato, M., Komatsu, R., Yamazaki, F., Kimura, I.,

Tsuchiya, S., Kori, H., Seo, K., et al. (2011). Circadian regulation of intracellular

G-protein signalling mediates intercellular synchrony and rhythmicity in the suprachiasmatic nucleus. Nat. Commun. 2, 327. https://doi.org/10.1038/

ncomms1316.

Finlin, B.S., Shao, H., Kadono-Okuda, K., Guo, N., and Andres, D.A. (2000).

Rem2, a new member of the Rem/Rad/Gem/Kir family of Ras-related

GTPases. Biochem. J. 347, 223–231. https://doi.org/10.1042/bj3470223.

Ginty, D.D., Kornhauser, J.M., Thompson, M.A., Bading, H., Mayo, K.E., Takahashi, J.S., and Greenberg, M.E. (1993). Regulation of CREB phosphorylation

in the suprachiasmatic nucleus by light and a circadian clock. Science 260,

238–241. https://doi.org/10.1126/science.8097062.

Gunton, J.E., Sisavanh, M., Stokes, R.A., Satin, J., Satin, L.S., Zhang, M., Liu,

S.M., Cai, W., Cheng, K., Cooney, G.J., et al. (2012). Mice deficient in GEM

GTPase show abnormal glucose homeostasis due to defects in beta-cell calcium handling. PLoS One 7, e39462. https://doi.org/10.1371/journal.pone.

0039462.

Hastings, M.H., Maywood, E.S., and Brancaccio, M. (2019). The mammalian

circadian timing system and the suprachiasmatic nucleus as its pacemaker.

Biology (Basel) 8, 13. https://doi.org/10.3390/biology8010013.

Herzog, E.D., Hermanstyne, T., Smyllie, N.J., and Hastings, M.H. (2017). Regulating the suprachiasmatic nucleus (SCN) circadian clockwork: interplay between cell-autonomous and circuit-level mechanisms. Cold Spring Harb. Perspect. Biol. 9, a027706. https://doi.org/10.1101/cshperspect.a027706.

Irwin, R.P., and Allen, C.N. (2007). Calcium response to retinohypothalamic

tract synaptic transmission in suprachiasmatic nucleus neurons. J. Neurosci.

27, 11748–11757. https://doi.org/10.1523/jneurosci.1840-07.2007.

Jagannath, A., Butler, R., Godinho, S.I.H., Couch, Y., Brown, L.A., Vasudevan,

S.R., Flanagan, K.C., Anthony, D., Churchill, G.C., Wood, M.J.A., et al. (2013).

The CRTC1-SIK1 pathway regulates entrainment of the circadian clock. Cell

154, 1100–1111. https://doi.org/10.1016/j.cell.2013.08.004.

Kim, D.Y., Choi, H.J., Kim, J.S., Kim, Y.S., Jeong, D.U., Shin, H.C., Kim, M.J.,

Han, H.C., Hong, S.K., and Kim, Y.I. (2005). Voltage-gated calcium channels

play crucial roles in the glutamate-induced phase shifts of the rat suprachiasmatic circadian clock. Eur. J. Neurosci. 21, 1215–1222. https://doi.org/10.

1111/j.1460-9568.2005.03950.x.

Kim, Y.I., and Dudek, F.E. (1991). Intracellular electrophysiological study of suprachiasmatic nucleus neurons in rodents: excitatory synaptic mechanisms.

J. Physiol. 444, 269–287. https://doi.org/10.1113/jphysiol.1991.sp018877.

Krey, J.F., Pasxca, S.P., Shcheglovitov, A., Yazawa, M., Schwemberger, R.,

Rasmusson, R., and Dolmetsch, R.E. (2013). Timothy syndrome is associated

with activity-dependent dendritic retraction in rodent and human neurons. Nat.

Neurosci. 16, 201–209. https://doi.org/10.1038/nn.3307.

OPEN ACCESS

zation of the suprachiasmatic nucleus under constant light conditions. Neuroscience 461, 1–10. https://doi.org/10.1016/j.neuroscience.2021.02.016.

Mizoro, Y., Yamaguchi, Y., Kitazawa, R., Yamada, H., Matsuo, M., Fustin, J.M.,

Doi, M., and Okamura, H. (2010). Activation of AMPA receptors in the suprachiasmatic nucleus phase-shifts the mouse circadian clock in vivo and in vitro.

PLoS One 5, e10951. https://doi.org/10.1371/journal.pone.0010951.

Murata, M., Cingolani, E., McDonald, A.D., Donahue, J.K., and Marba´n, E.

(2004). Creation of a genetic calcium channel blocker by targeted gem gene

transfer in the heart. Circ. Res. 95, 398–405. https://doi.org/10.1161/01.RES.

0000138449.85324.c5.

Nanou, E., and Catterall, W.A. (2018). Calcium channels, synaptic plasticity,

and neuropsychiatric disease. Neuron 98, 466–481. https://doi.org/10.1016/

j.neuron.2018.03.017.

Nelson, D.E., and Takahashi, J.S. (1991). Sensitivity and integration in a visual

pathway for circadian entrainment in the hamster (Mesocricetus auratus).

J. Physiol. 439, 115–145. https://doi.org/10.1113/jphysiol.1991.sp018660.

Obrietan, K., Impey, S., and Storm, D.R. (1998). Light and circadian rhythmicity

regulate MAP kinase activation in the suprachiasmatic nuclei. Nat. Neurosci. 1,

693–700. https://doi.org/10.1038/3695.

Okamura, H. (2007). Suprachiasmatic nucleus clock time in the mammalian

circadian system. Cold Spring Harb Symp. Quant. Biol. 72, 551–556. https://

doi.org/10.1101/sqb.2007.72.033.

Scamps, F., Sangari, S., Bowerman, M., Rousset, M., Bellis, M., Cens, T., and

Charnet, P. (2015). Nerve injury induces a Gem-GTPase-dependent downregulation of P/Q-type Ca2+ channels contributing to neurite plasticity in dorsal

root ganglion neurons. Pflugers Arch. 467, 351–366. https://doi.org/10.1007/

s00424-014-1520-4.

Schmutz, I., Chavan, R., Ripperger, J.A., Maywood, E.S., Langwieser, N., Jurik, A., Stauffer, A., Delorme, J.E., Moosmang, S., Hastings, M.H., et al. (2014).

A specific role for the REV-ERBa-controlled L-type voltage-gated calcium

channel CaV1.2 in resetting the circadian clock in the late night. J. Biol.

Rhythms 29, 288–298. https://doi.org/10.1177/0748730414540453.

Schwartz, W.J., Carpino, A., Jr., de la Iglesia, H.O., Baler, R., Klein, D.C., Nakabeppu, Y., and Aronin, N. (2000). Differential regulation of fos family genes in

the ventrolateral and dorsomedial subdivisions of the rat suprachiasmatic nucleus. Neuroscience 98, 535–547. https://doi.org/10.1016/s0306-4522(00)

00140-8.

Schwartz, W.J., and Zimmerman, P. (1990). Circadian timekeeping in BALB/c

and C57BL/6 inbred mouse strains. J. Neurosci. 10, 3685–3694. https://doi.

org/10.1523/jneurosci.10-11-03685.1990.

LeGates, T.A., Fernandez, D.C., and Hattar, S. (2014). Light as a central modulator of circadian rhythms, sleep and affect. Nat. Rev. Neurosci. 15, 443–454.

https://doi.org/10.1038/nrn3743.

Shigeyoshi, Y., Taguchi, K., Yamamoto, S., Takekida, S., Yan, L., Tei, H., Moriya, T., Shibata, S., Loros, J.J., Dunlap, J.C., and Okamura, H. (1997). Lightinduced resetting of a mammalian circadian clock is associated with rapid induction of the mPer1 transcript. Cell 91, 1043–1053. https://doi.org/10.1016/

s0092-8674(00)80494-8.

Liput, D.J., Lu, V.B., Davis, M.I., Puhl, H.L., and Ikeda, S.R. (2016). Rem2, a

member of the RGK family of small GTPases, is enriched in nuclei of the basal

ganglia. Sci. Rep. 6, 25137. https://doi.org/10.1038/srep25137.

Takahashi, J.S. (2017). Transcriptional architecture of the mammalian circadian clock. Nat. Rev. Genet. 18, 164–179. https://doi.org/10.1038/nrg.2016.

150.

Maguire, J., Santoro, T., Jensen, P., Siebenlist, U., Yewdell, J., and Kelly, K.

(1994). Gem: an induced, immediate early protein belonging to the Ras family.

Science 265, 241–244. https://doi.org/10.1126/science.7912851.

Ward, Y., Yap, S.F., Ravichandran, V., Matsumura, F., Ito, M., Spinelli, B., and

Kelly, K. (2002). The GTP binding proteins Gem and Rad are negative regulators of the Rho–Rho kinase pathway. J. Cell Biol. 157, 291–302. https://doi.

org/10.1083/jcb.200111026.

Magyar, J., Kiper, C.E., Sievert, G., Cai, W., Shi, G.X., Crump, S.M., Li, L., Niederer, S., Smith, N., Andres, D.A., and Satin, J. (2012). Rem-GTPase regulates

cardiac myocyte L-type calcium current. Channels (Austin) 6, 166–173. https://

doi.org/10.4161/chan.20192.

Manning, J.R., Yin, G., Kaminski, C.N., Magyar, J., Feng, H.Z., Penn, J., Sievert, G., Thompson, K., Jin, J.P., Andres, D.A., and Satin, J. (2013). Rad

GTPase deletion increases L-type calcium channel current leading to

increased cardiac contraction. J. Am. Heart Assoc. 2, e000459. https://doi.

org/10.1161/jaha.113.000459.

Matsuo, M., Seo, K., Mizuguchi, N., Yamazaki, F., Urabe, S., Yamada, N., Doi,

M., Tominaga, K., and Okamura, H. (2021). Role of a2d3 in cellular synchroni-

Yagi, T., Tokunaga, T., Furuta, Y., Nada, S., Yoshida, M., Tsukada, T., Saga, Y.,

Takeda, N., Ikawa, Y., and Aizawa, S. (1993). A novel ES cell line, TT2, with high

germline-differentiating potency. Anal. Biochem. 214, 70–76. https://doi.org/

10.1006/abio.1993.1458.

Yamaguchi, S., Mitsui, S., Miyake, S., Yan, L., Onishi, H., Yagita, K., Suzuki, M.,

Shibata, S., Kobayashi, M., and Okamura, H. (2000). The 5’ upstream region of

mPer1 gene contains two promoters and is responsible for circadian oscillation. Curr. Biol. 10, 873–876. https://doi.org/10.1016/s0960-9822(00)00602-3.

Yamaguchi, S., Isejima, H., Matsuo, T., Okura, R., Yagita, K., Kobayashi, M.,

and Okamura, H. (2003). Synchronization of cellular clocks in the

Cell Reports 39, 110844, May 24, 2022 9

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Report

suprachiasmatic nucleus. Science 302, 1408–1412. https://doi.org/10.1126/

science.1089287.

Zamponi, G.W. (2016). Targeting voltage-gated calcium channels in neurological and psychiatric diseases. Nat. Rev. Drug Discov. 15, 19–34. https://doi.

org/10.1038/nrd.2015.5.

Yamaguchi, Y., Suzuki, T., Mizoro, Y., Kori, H., Okada, K., Chen, Y., Fustin,

J.M., Yamazaki, F., Mizuguchi, N., Zhang, J., et al. (2013). Mice genetically

deficient in vasopressin V1a and V1b receptors are resistant to jet lag. Science

342, 85–90. https://doi.org/10.1126/science.1238599.

Zhang, J., Chang, L., Chen, C., Zhang, M., Luo, Y., Hamblin, M., Villacorta, L.,

Xiong, J.W., Chen, Y.E., Zhang, J., and Zhu, X. (2011). Rad GTPase inhibits

cardiac fibrosis through connective tissue growth factor. Cardiovasc. Res.

91, 90–98. https://doi.org/10.1093/cvr/cvr068.

10 Cell Reports 39, 110844, May 24, 2022

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Report

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE

SOURCE

IDENTIFIER

Antibodies

Phospho-p44/42 MAPK (Erk1/2) (Thr202/Tyr204) antibody

Cell ignaling Tech

Cat#9101

Gem monoclonal antibody (M01), clone 4B12

Abnova

Cat#H00002669-M01

c-Fos antibody

Abcam

Cat#ab7963

Chemicals, peptides, and recombinant proteins

Toluidine blue

Sigma-Aldrich

Cat#T3260

TRIzol reagent

Thermo

Cat#15596026

RNeasy micro kit

Qiagen

Cat#74004

Tetrodotoxin

Wako

Cat# 206-11071

One-Cycle Target Labelling and Control

Reagents Kit

Affymetrix

Discontinued

3,3’-Diaminobenzidine (DAB)

Wako

Cat#R-074N

Goat Biotinylated anti-rabbit IgG

Vector lab

Cat# BP-9100-50

Lipofectamine 2000

Thermo

Cat#11668027

Bicculine

Wako

Cat#0130

proteinase K

Wako

Cat#161-28701

RNAse A

Wako

Cat#318-06391

Avidin/biotinylated horseradish peroxidase

Vector lab

Cat#PK-4005

Diaminobenzidine

Wako

Cat#349-00903

Entellan

Merck

Cat#107960

SuperScript III First-Strand Synthesis SuperMix

Thermo

Cat#18080400

Platinum SYBR Green qPCR SuperMix-UDG

Thermo

Cat#11733038

PrimeScript RT reagent kits

Takara

Cat#RR037A

Thunderbird SYBR qPCR mix

Toyobo

Cat#QPS-201

2-mercaptoethanol

Nacalai

Cat#21417-52

RNeasy Mini Kit

Qiagen

Cat#74104

SuperScriptTM VILOTM cDNA Synthesis

Thermo

Cat#11754050

ROCHE Digoxigenin-11-UTP

MilliporeSigma

Cat#12352200

Nifedipine

Wako

Cat# 141-05783

Promega

E1910

Mouse: Gem/

In this paper

N/A

Mouse: Per1-promoter-luc

Curr Biol. 10(14):873-6,2000.

N/A

pGL3-Promoter

Promega

E1751

pRL-CMV Renilla reporter

Promega

E2261

GraphPad

https://www.graphpad.com/

Critical commercial assays

Dual-Luciferase Reporter Assay System

Experimental models: Organisms/strains

Recombinant DNA

Software and algorithms

Prism 5

RESOURCE AVAILABILITY

Lead contact

Further information and requests for reagents should be directed to and will be fulfilled by Hitoshi Okamura (e-mail: okamura.hitoshi.

4u@kyoto-u.ac.jp).

Cell Reports 39, 110844, May 24, 2022 e1

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Report

Materials availability

All unique/stable reagents generated in this study are available from the lead contact without restriction.

Data and code availability

d All data reported in this paper will be shared by the lead contact upon request.

d This paper does not report original code.

d Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.

EXPERIMENTAL MODEL AND SUBJECT DETAILS

We developed and mated male and female C57BL6 backcrossed Gem knockout mice (Gem/) and Gem+/+ mice in housing conditions under 12-h light: 12-h dark (LD) cycles at 22 ± 2 C, with food and water provided ad libitum. We used male mice to avoid the

effect of sexual cycle on circadian rhythms. At 8–12 weeks of age, mice were transferred to individually housed conditions under 12-h

light: 12-h dark (LD) cycles and kept there for at least one week before the behavioral analysis. For electrophysiological studies, adult

Gem/ and Gem+/+ mice (6- to 10-week-old) were housed in the same condition for at least 2 weeks prior to experiments. For organotypic slice culture study, we sampled SCN from both sexes of neonatal (4- to 7- day old) Gem/ and Gem+/+ mice harboring

Per1-luc transgene, and luciferase experiments were performed after 2 to 4 weeks of initiation of culture.

All animal experimental procedures were pre-approved by the Animal Experimentation Committee at Kyoto University and the

Institutional Animal Care and Use Committee of RIKEN Kobe Branch. Experimental protocols for mice were in accordance with

the guidelines for animal experiments approved by the Animal Ethics Committee at Kyoto University and RIKEN Regulations for

the Animal Experiments.

METHOD DETAILS

Screening strategy of SCN gene project

We first screened the involvement of RGK protein expressions in the SCN by in situ hybridization at various circadian time point by a

screening strategy called SCN-Gene Project (Okamura, 2007), in which we already found key signaling molecules for the circadian

generation and synchronization of cellular clocks in the SCN (Doi et al., 2011; Yamaguchi et al., 2013,; Matsuo et al., 2021). In the

project, we (i) used histochemistry to identify genes whose expression is enriched in the mouse SCN, (ii) generated mutant mice lacking candidate genes of interest, and (iii) measured the locomotor activity at various environmental conditions.

Generation of gem/ mice

The Gem knockout mice (Accession No. CDB0673K: http://www2.clst.riken.jp/arg/mutant%20mice%20list.html) were generated. The

targeting construct, containing a part of the first exon and full range of the second exon of Gem, was prepared using genomic DNA fragments from a BAC clone. Briefly, 7.0 kb adjacent upstream of Gem and 3.4 kb downstream of Gem was amplified by PCR and sub-cloned

into PGK-Neo-pA/DT-ApA vector and verified by full sequencing. The targeting constructs were linearized with NotI and electroporated

into the TT2 embryonic stem (ES) cells (Yagi et al., 1993). The construct was introduced into C57BL/6/CBA background ES cells, and two

independent chimeric mouse lines were established. These lines were independently crossed with C57BL/6 mice to produce F1N0 heterozygous mice, and then crossed further for at least three times before use in experiments. Gem/ mice did not have gross abnormalities. They were fertile and were born at the expected Mendelian ratio. Disruption of Gem in the genome was confirmed by Southern blotting, conventional PCR, and in situ hybridization. We further confirmed targeting of Gem by western blotting with micro-punched SCN

extracts.

Genotypes were determined by Southern blot and PCR. NheI-digested DNA was Southern blotted and hybridized with a 32Plabeled external probe. Genotyping primers we used for detecting WT and KO alleles were 50 -CCGTGCATTGGCTTTATC-TT-30

(Fw primer for WT), 50 -TCGCCTTCTTGACGAGTTCT-30 (Fw primer for KO), and 50 -AGCTCAGGCCTCCTAAGTCC-30 (common

Rev primer for WT and KO).

Gem activates Rho kinase-mediated cytoskeletal reorganization, such as stress fiber formation and neurite retraction (Krey et al.,

2013; Ward et al., 2002). Gem-knockout mice (Gem/) develop normally at young ages. We also assessed behaviors at 8–10weeks-of-age and found no evidence of significant neurological impairment such as cerebellar ataxia, vestibular ataxia, or chorea.

Examination of animal behavior

To determine the impact of Gem deficiency on the circadian clock system, we monitored the circadian activity of mice and their response

to a brief light pulse at 8-12-weeks-of-age. Gem/ mice and their Gem+/+ littermates were individually housed in circadian activity-monitoring chambers. Food and water were given ad libitum and daily activity of each mouse was recorded using infrared sensors. Locomotor

activity was detected with passive (pyroelectric) infrared sensors (FA-05 F5B; Omron), and the data obtained were analyzed with Clocklab software (Actimetrics) developed on MatLab (Mathworks). Mice were initially maintained on 12 h light/12 h dark (LD) cycles for at least

10 days to establish entrainment before being transferred to constant darkness (DD). A free-running period was determined with a linear

regression line fit to the activity onset of 14 consecutive days in DD.

e2 Cell Reports 39, 110844, May 24, 2022

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

Report

OPEN ACCESS

For the phase-dependent phase-shift experiments, mice in DD were exposed to a 30 min light pulse (200 lux) at eight different

designated time points: CT2, 5, 8, 11, 14, 17, 20, and 23. The effect of light intensity on phase-shift was examined by changing light

intensities (2, 30, 200, and 1000 lux) at CT14. Phase-shifts were quantified as the time difference between regression lines of activity

onset before and after light application. Regression lines were made based on the activity onset 7 days before light stimulation, and

7 days activity onset starting 3 days after the light stimulation.

All animal procedures described in this study were approved by the Animal Research Committee of Kyoto University (2010-43) and

the Committee for Animal Research of Kobe University (P060601).

Electrophysiological recordings

For electrophysiological studies, adult mice (6- to 10-week-old) were entrained for at least 2 weeks on a 12 h light/dark schedule prior

to experiments. SCN slices (250 mm-thick) were prepared with a Microslicer (DTK-300K, Dohsaka EM, Japan) in ice-cold reduced Ca2+

artificial cerebrospinal fluid (aCSF) containing (in mM): 120.0 NaCl, 2.5 KCl, 1.2 NaH2PO4, 10.0 glucose, 26.0 NaHCO3, 5.0 MgCl2, and

0.5 CaCl2 (Chuhma et al., 2001). Whole-cell patch-clamp recordings in voltage-clamp mode from SCN neurons were carried out. SCN

neurons were randomly selected. Patch pipettes were prepared from thick-walled borosilicate glass to a tip resistance of 5–8 MU. The

intracellular recording medium contained (in mM): 155.0 Cs-methanesulfonate, 5.0 NaCl, 10.0 HEPES, 0.2 ethylenediaminetetraacetic

acid (EGTA), and 3.0 MgCl2. Corrections were made for the liquid junction potential (18 mV). To reduce large Na+ currents and Cl

currents in SCN neurons, tetrodotoxin (TTX, 1 mM) and bicuculline (40 mM) were added to the recording medium. The membrane potential was held at 80 mV, and voltage steps were applied between 90 and +70 mV per 10 mV step. Active currents were generated

from approximately 20 mV. Current is normalized to cell membrane capacitance to produce current density for each cell (pA/pF).

Luciferase activity

Organotypic SCN slice cultures from Per1-luc neonatal transgenic mice (4- to 7-day-old) were obtained as described previously (Yamaguchi et al., 2003, 2000). Gem+/ mice were crossed to Per1-luc transgenic mice to generate Per1-luc transgenic mice harboring either a

Gem/ or Gem+/+ genotype. Our bioluminescent system is able to determine the phase of the SCN with a high time-resolution (5 min). To

align the phase condition between slices, tissues were entrained with a 36 h-bath application of 10 mg/mL cycloheximide (CHX) (Yamaguchi et al., 2003), and the phase of the circadian clock in each SCN was determined by the peak time of Per1-luciferase activity. SCN

slice cultures were maintained in a sealed 24-well cell culture plate; during bioluminescence recording, each well contained 240 mL culture medium with 1 mM luciferin at 35 C. For high K+ (50 mM) stimulation, SCN slice cultures were transferred 6 h after the first peak to

control medium (50% minimum essential medium, 50% Hank’s balanced salt solution, 36 mM glucose, and penicillin/streptomycin),

with or without K+ (50 mM) and/or nifedipine (20 mM), for 30 min at 35 C. The tissues were then washed three times with control medium

for 30 min at 35 C, and returned to the original culture medium. Sample sizes were vehicle-control (n = 10), high K+ (n = 11), high K+ plus

nifedipine (n = 11), and nifedipine (n = 10), respectively. A one-way ANOVA was used to analyze these data.

Radioisotopic in situ hybridization

In situ hybridization was performed as previously described (Shigeyoshi et al., 1997). Briefly, paraformaldehyde-fixed brains were

frozen and cryoprotected. Serial coronal sections (40 mm-thick) were prepared from the rostral end to the caudal end of the SCN

with a cryostat. Free-floating tissue sections were then transferred through 4 3 saline sodium citrate (SSC) buffer, proteinase K

(1 mg/mL) in 0.1 M Tris buffer [pH 8.0], 50 mM EDTA for 15 min at 37 C, 0.25% acetic anhydride in 0.1 M triethanolamine for

10 min, and 4 3 SSC for 10 min. The sections were then incubated in hybridization buffer [55% formamide, 10% dextran sulfate,

10 mM Tris-HCl (pH 8.0), 1 mM EDTA (pH 8.0), 0.6 M NaCl, 0.2% N-laurylsarcosine, 500 mg/mL tRNA, 1 3 Denhardt’s, 0.25%

SDS, and 10 mM dithiothreitol (DTT)] containing radiolabeled riboprobes for 16 h at 60 C.

Radiolabeled molecular-specific riboprobes were prepared based on the following sequences: for Gem, nucleotides 254–988 of

Gem (NM_010276.4); for Rad, nucleotides 1–543 of (NM_019662.1); for Rem1, nucleotides 100–629 of (NM_009047.4); for Rem2, nucleotides 103–700 of (NM_080726.2); for Cacna1s, nucleotides 2291–2851 of (XM_358335.4); for Cacna1c, nucleotides 5470–5914 of

(NM_009781.3); for Cacna1d, nucleotides 5504–5979 of (NM_028981.2); for Cacna1f, nucleotides 4837–5261 of (NM_019582.2); for

Cacna1a, nucleotides 3339–3736 of (NM_007578.3); for Cacna1b, nucleotides 5910–6406 of (NM_007579.1); for Cacna1e, nucleotides 1353–1832 of (AK171983.1); for Cacna1g, nucleotides 7155–7608 of (NM_009783.1); for Cacna1h, nucleotides 6968–7397 of

(NM_021415.3); for Cacna1i, nucleotides 6874–7255 of (NM_001044308.2); for Cacna2d1, nucleotides 2141–2683 of

(NM_001110843.1); for Cacna2d2, nucleotides 1172–1643 of (NM_001174047.1); for Cacna2d3, nucleotides 2443–2986

of (NM_009785.1); for Cacnb1, nucleotides 193–748 of (NM_031173.3); for Cacnb2, nucleotides 1434–2003 of (NM_023116.4); for

Cacnb3, nucleotides 1699–2179 of (NM_007581.2); for Cacnb4, nucleotides 1278–1799 of (NM_001037099.1). The corresponding

cDNA fragment was cloned and used as a template for the generation of riboprobes. The riboprobes were radiolabeled with [33P]

UTP (PerkinElmer, Waltham, MA), using a standard protocol for the cRNA synthesis. Following a high-stringency post-hybridization

wash, the sections were treated with RNase A. Air-dried sections were exposed to X-ray film (Kodak Biomax).

Autoradiographic films (Kodak, Biomax) were quantified with MCID imaging software (Imaging Research Inc., Canada) after conversion into the relative optical densities using 14C-autoradiographic microscales (Amersham, UK); ten SCN sections were then

summed.

Cell Reports 39, 110844, May 24, 2022 e3

A Self-archived copy in

Kyoto University Research Information Repository

https://repository.kulib.kyoto-u.ac.jp

ll

OPEN ACCESS

Report

In situ hybridization of gem with digoxigenin-labeled riboprobes

Digoxigenin-labeled probes allow for a better resolution than isotope probes for analyzing the cellular distribution of mRNA. We,

therefore, prepared digoxigenin-labeled antisense cRNA probes using digoxigenin-UTP (Roche Diagnostics) following a standard

protocol for cRNA synthesis. Tissue preparation, pre-hybridization, hybridization, and post-hybridization washing were identical

as in isotope probe hybridization, with the exception that we used 20 mm-thick sections of the SCN. Brain sections hybridized

with the digoxigenin-labeled probe were processed for immunochemistry with a nucleic acid detection kit (Roche Diagnostics). Signals were visualized in a solution containing nitroblue tetrazolium salt (0.34 mg/mL) and 5-bromo-4-chloro-3-indolyl phosphate toluidinium salt (0.18 mg/mL) (Roche Diagnostics).

Immunohistochemistry for pERK and cFos

For immunohistochemical labeling of pERK and cFos, animals were anesthetized and systemically perfused with 25 mL of cold fixative containing 4% paraformaldehyde in 0.1 M phosphate buffer (PB) (Doi et al., 2011). Isolated brains were transferred to the same

fixative for 12 h at 4 C, and then immersed in 20% sucrose in 0.1 M PB for cryoprotection. Coronal brain cryosections (20 mm-thick)

were processed for free-floating immunohistochemistry. We used a rabbit polyclonal antibody specific for the phosphorylated forms

of ERK 1 and 2 (Cell Signaling; catalog code 9101) and a rabbit antibody for cFos protein identification (Abcam; catalog code 7963).

Free-floating sections were pretreated with hydrogen peroxide (1.5% in 0.1 M PB, for 20 min at 4 C) and blocked with 5% horse

serum (in 0.1 M PB) for 1 h at room temperature. The sections were then incubated with pERK antibody [1:500 dilution, in 0.1 M

PB containing 0.3% Triton X-100 (PBX)], or with cFos antibody [1:8000] for 12 h at room temperature. After washing with PBX,

the sections were incubated with a biotinylated anti-rabbit IgG secondary antibody (1:500 dilution in PBX) (Vector Laboratories)

for 1 h at room temperature. The tissues were then subjected to standard avidin-biotin-immunoperoxidase staining (Vectorstain Elite

ABC kit, Vector Laboratories). Immunoreactivity was visualized with 3,30 -diaminobenzidine (DAB). Immunostained sections were

then washed with 50 mM Tris-HCl buffer (pH 7.5), dehydrated in ethanol, and coverslipped with Entellan mounting medi ...

参考文献をもっと見る

全国の大学の
卒論・修論・学位論文

一発検索!

この論文の関連論文を見る