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 「Thermodynamic and kinetic stabilities of transmembrane helix bundles as revealed by single-pair FRET analysis: Effects of the number of membrane-spanning segments and cholesterol」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

Thermodynamic and kinetic stabilities of transmembrane helix bundles as revealed by single-pair FRET analysis: Effects of the number of membrane-spanning segments and cholesterol

Yano, Yoshiaki Watanabe, Yuta Matsuzaki, Katsumi 京都大学 DOI:10.1016/j.bbamem.2020.183532

2021.03.01

概要

The tertiary structures and conformational dynamics of transmembrane (TM) helical proteins are maintained by the interhelical interaction network in membranes, although it is complicated to analyze the underlying driving forces because the amino acid sequences can involve multiple and various types of interactions. To obtain insights into basal and common effects of the number of membrane-spanning segments and membrane cholesterol, we measured stabilities of helix bundles composed of simple TM helices (AALALAA)3 (1TM) and (AALALAA)3-G5-(AALALAA)3 (2TM). Association–dissociation dynamics for 1TM–1TM, 1TM–2TM, and 2TM–2TM pairs were monitored to compare stabilities of 2-, 3-, and 4-helical bundles, respectively, with single-pair fluorescence resonance energy transfer (sp-FRET) in liposome membranes. Both thermodynamic and kinetic stabilities of the helix bundles increased with a greater number of membrane-spanning segments in POPC. The presence of 30 mol% cholesterol strongly enhanced the formation of 1TM–1TM and 1TM–2TM bundles (~ − 9 kJ mol−1), whereas it only weakly stabilized the 2TM–2TM bundle (~ − 3 kJ mol−1). Fourier transform infrared-polarized attenuated total reflection (ATR-FTIR) spectroscopy revealed an ~30° tilt of the helix axis relative to bilayer normal for the 1TM–2TM pair in the presence of cholesterol, suggesting the formation of a tilted helix bundle to release high lateral pressure at the center of cholesterol-containing membranes. These results demonstrate that the number of membrane-spanning segments affects the stability and structure of the helix bundle, and their cholesterol-dependences. Such information is useful to understand the basics of folding and assembly of multispanning TM proteins.

この論文で使われている画像

参考文献

[1] G.G. Gregorio, M. Masureel, D. Hilger, D.S. Terry, M. Juette, H. Zhao, Z. Zhou, J.M. PerezAguilar, M. Hauge, S. Mathiasen, J.A. Javitch, H. Weinstein, B.K. Kobilka, S.C. Blanchard,

Single-molecule analysis of ligand efficacy in β2AR-G-protein activation, Nature, 547 (2017)

68–73.

[2] E.O. Grushevskyi, T. Kukaj, R. Schmauder, A. Bock, U. Zabel, T. Schwabe, K. Benndorf, M.J.

Lohse, Stepwise activation of a class C GPCR begins with millisecond dimer rearrangement,

Proc. Natl. Acad. Sci. U.S.A., 116 (2019) 10150–10155.

[3] Y. Kofuku, T. Ueda, J. Okude, Y. Shiraishi, K. Kondo, T. Mizumura, S. Suzuki, I. Shimada,

Functional dynamics of deuterated β2-adrenergic receptor in lipid bilayers revealed by NMR

spectroscopy, Angew. Chem. Int. Ed. Engl., 53 (2014) 13376–13379.

[4] M.J. Lohse, K.P. Hofmann, Spatial and Temporal Aspects of Signaling by G-Protein-Coupled

Receptors, Mol. Pharmacol., 88 (2015) 572–578.

[5] J.A. Hern, A.H. Baig, G.I. Mashanov, B. Birdsall, J.E. Corrie, S. Lazareno, J.E. Molloy, N.J.

Birdsall, Formation and dissociation of M1 muscarinic receptor dimers seen by total internal

reflection fluorescence imaging of single molecules, Proc. Natl. Acad. Sci. U.S.A., 107 (2010)

2693–2698.

[6] R.E. Jefferson, T.M. Blois, J.U. Bowie, Membrane proteins can have high kinetic stability, J.

Am. Chem. Soc., 135 (2013) 15183–15190.

[7] R.E. Jefferson, D. Min, K. Corin, J.Y. Wang, J.U. Bowie, Applications of Single-Molecule

Methods to Membrane Protein Folding Studies, J. Mol. Biol., 430 (2018) 424–437.

[8] R.S. Kasai, A. Kusumi, Single-molecule imaging revealed dynamic GPCR dimerization, Curr.

Opin. Cell Biol., 27 (2014) 78–86.

21

A Self-archived copy in

Kyoto University Research Information Repository

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

[9] R. Lamichhane, J.J. Liu, K.L. White, V. Katritch, R.C. Stevens, K. Wuthrich, D.P. Millar,

Biased Signaling of the G-Protein-Coupled Receptor β2AR Is Governed by Conformational

Exchange Kinetics, Structure, 28 (2020) 371–377.

[10] R. Maeda, T. Sato, K. Okamoto, M. Yanagawa, Y. Sako, Lipid-Protein Interplay in

Dimerization of Juxtamembrane Domains of Epidermal Growth Factor Receptor, Biophys. J.,

114 (2018) 893–903.

[11] F. Cornelius, Modulation of Na,K-ATPase and Na-ATPase activity by phospholipids and

cholesterol. I. Steady-state kinetics, Biochemistry, 40 (2001) 8842–8851.

[12] A.G. Lee, How lipids affect the activities of integral membrane proteins, Biochim. Biophys.

Acta, 1666 (2004) 62–87.

[13] M. Stangl, D. Schneider, Functional competition within a membrane: Lipid recognition vs.

transmembrane helix oligomerization, Biochim. Biophys. Acta, 1848 (2015) 1886–1896.

[14] E. van den Brink-van der Laan, J.A. Killian, B. de Kruijff, Nonbilayer lipids affect peripheral

and integral membrane proteins via changes in the lateral pressure profile, Biochim. Biophys.

Acta, 1666 (2004) 275–288.

[15] V. Anbazhagan, D. Schneider, The membrane environment modulates self-association of the

human GpA TM domain--implications for membrane protein folding and transmembrane

signaling, Biochim. Biophys. Acta, 1798 (2010) 1899–1907.

[16] H. Hong, T.M. Blois, Z. Cao, J.U. Bowie, Method to measure strong protein–protein

interactions in lipid bilayers using a steric trap, Proc. Natl. Acad. Sci. U.S.A., 107 (2010)

19802–19807.

[17] H. Hong, J.U. Bowie, Dramatic destabilization of transmembrane helix interactions by features

of natural membrane environments, J. Am. Chem. Soc., 133 (2011) 11389–11398.

22

A Self-archived copy in

Kyoto University Research Information Repository

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

[18] S. Mall, R. Broadbridge, R.P. Sharma, J.M. East, A.G. Lee, Self-association of model

transmembrane α-helices is modulated by lipid structure, Biochemistry, 40 (2001) 12379–

12386.

[19] A. Nash, R. Notman, A.M. Dixon, De novo design of transmembrane helix–helix interactions

and measurement of stability in a biological membrane, Biochim. Biophys. Acta, 1848 (2015)

1248–1257.

[20] Y. Song, E.J. Hustedt, S. Brandon, C.R. Sanders, Competition between homodimerization and

cholesterol binding to the C99 domain of the amyloid precursor protein, Biochemistry, 52

(2013) 5051–5064.

[21] E. Sparr, W.L. Ash, P.V. Nazarov, D.T. Rijkers, M.A. Hemminga, D.P. Tieleman, J.A. Killian,

Self-association of transmembrane α-helices in model membranes: importance of helix

orientation and role of hydrophobic mismatch, J. Biol. Chem., 280 (2005) 39324–39331.

[22] M.G. Teese, D. Langosch, Role of GxxxG Motifs in Transmembrane Domain Interactions,

Biochemistry, 54 (2015) 5125–5135.

[23] Y. Yano, K. Kondo, R. Kitani, A. Yamamoto, K. Matsuzaki, Cholesterol-induced lipophobic

interaction between transmembrane helices using ensemble and single-molecule fluorescence

resonance energy transfer, Biochemistry, 54 (2015) 1371–1379.

[24] Y. Yano, K. Matsuzaki, Measurement of thermodynamic parameters for hydrophobic

mismatch 1: self-association of a transmembrane helix, Biochemistry, 45 (2006) 3370–3378.

[25] Y. Yano, A. Yamamoto, M. Ogura, K. Matsuzaki, Thermodynamics of insertion and selfassociation of a transmembrane helix: a lipophobic interaction by phosphatidylethanolamine,

Biochemistry, 50 (2011) 6806–6814.

[26] W.P. Russ, D.M. Engelman, The GxxxG motif: a framework for transmembrane helix-helix

association, J. Mol. Biol., 296 (2000) 911–919.

23

A Self-archived copy in

Kyoto University Research Information Repository

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

[27] R.F. Walters, W.F. DeGrado, Helix-packing motifs in membrane proteins, Proc. Natl. Acad.

Sci. U.S.A., 103 (2006) 13658–13663.

[28] S.Q. Zhang, D.W. Kulp, C.A. Schramm, M. Mravic, I. Samish, W.F. DeGrado, The

membrane- and soluble-protein helix-helix interactome: similar geometry via different

interactions, Structure, 23 (2015) 527–541.

[29] K. Yamamoto, U.H. Durr, J. Xu, S.C. Im, L. Waskell, A. Ramamoorthy, Dynamic interaction

between membrane-bound full-length cytochrome P450 and cytochrome b5 observed by solidstate NMR spectroscopy, Sci. Rep., 3 (2013) 2538.

[30] K. Yamamoto, M. Gildenberg, S. Ahuja, S.C. Im, P. Pearcy, L. Waskell, A. Ramamoorthy,

Probing the transmembrane structure and topology of microsomal cytochrome-p450 by solidstate NMR on temperature-resistant bicelles, Sci. Rep., 3 (2013) 2556.

[31] Y. Yano, K. Kondo, Y. Watanabe, T.O. Zhang, J.J. Ho, S. Oishi, N. Fujii, M.T. Zanni, K.

Matsuzaki, GXXXG-Mediated Parallel and Antiparallel Dimerization of Transmembrane

Helices and Its Inhibition by Cholesterol: Single-Pair FRET and 2D IR Studies, Angew. Chem.

Int. Ed. Engl., 56 (2017) 1756–1759.

[32] N. Castillo, L. Monticelli, J. Barnoud, D.P. Tieleman, Free energy of WALP23 dimer

association in DMPC, DPPC, and DOPC bilayers, Chem. Phys. Lipids, 169 (2013) 95–105.

[33] M.N. Nishizawa, K., Sequence-nonspecific stabilization of transmembrane helical peptide

dimer in lipid raft-like bilayers in atomistic simulations. I. Dimerization free energy and impact

of lipid–peptide potential energy, Ann. Biomed. Res., 1 (2018) 105.

[34] A.B. Pawar, S.A. Deshpande, S.M. Gopal, T.A. Wassenaar, C.A. Athale, D. Sengupta,

Thermodynamic and kinetic characterization of transmembrane helix association, Phys. Chem.

Chem. Phys., 17 (2015) 1390–1398.

[35] G.R. Bartlett, Phosphorus assay in column chromatography, J. Biol. Chem., 234 (1959) 466–

468.

24

A Self-archived copy in

Kyoto University Research Information Repository

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

[36] Y. Yano, T. Takemoto, S. Kobayashi, H. Yasui, H. Sakurai, W. Ohashi, M. Niwa, S. Futaki, Y.

Sugiura, K. Matsuzaki, Topological stability and self-association of a completely hydrophobic

model transmembrane helix in lipid bilayers, Biochemistry, 41 (2002) 3073–3080.

[37] C. Joo, T. Ha, Preparing sample chambers for single-molecule FRET, Cold Spring Harb.

Protoc., 2012 (2012) 1104–1108.

[38] S.A. McKinney, C. Joo, T. Ha, Analysis of single-molecule FRET trajectories using hidden

Markov modeling, Biophys. J., 91 (2006) 1941–1951.

[39] S. Ramadurai, A. Holt, V. Krasnikov, G. van den Bogaart, J.A. Killian, B. Poolman, Lateral

diffusion of membrane proteins, J. Am. Chem. Soc., 131 (2009) 12650–12656.

[40] L.M. Smith, J.L. Rubenstein, J.W. Parce, H.M. McConnell, Lateral diffusion of M-13 coat

protein in mixtures of phosphatidylcholine and cholesterol, Biochemistry, 19 (1980) 5907–

5911.

[41] D.W. Tank, E.S. Wu, P.R. Meers, W.W. Webb, Lateral diffusion of gramicidin C in

phospholipid multibilayers. Effects of cholesterol and high gramicidin concentration, Biophys.

J., 40 (1982) 129–135.

[42] E.S. Wu, C.S. Yang, Lateral diffusion of cytochrome P-450 in phospholipid bilayers,

Biochemistry, 23 (1984) 28–33.

[43] F.A. Nezil, M. Bloom, Combined influence of cholesterol and synthetic amphiphillic peptides

upon bilayer thickness in model membranes, Biophys. J., 61 (1992) 1176–1183.

[44] H. Yang, S. Yang, J. Kong, A. Dong, S. Yu, Obtaining information about protein secondary

structures in aqueous solution using Fourier transform IR spectroscopy, Nat. Protoc., 10 (2015)

382–396.

[45] Y.B. Yu, P.L. Privalov, R.S. Hodges, Contribution of translational and rotational motions to

molecular association in aqueous solution, Biophys. J., 81 (2001) 1632–1642.

25

A Self-archived copy in

Kyoto University Research Information Repository

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

[46] H. Heerklotz, A. Tsamaloukas, Gradual change or phase transition: characterizing fluid lipidcholesterol membranes on the basis of thermal volume changes, Biophys. J., 91 (2006) 600–

607.

[47] R.F. Epand, A. Ramamoorthy, R.M. Epand, Membrane lipid composition and the interaction of

pardaxin: the role of cholesterol, Protein Pept. Lett., 13 (2006) 1–5.

[48] A. Ramamoorthy, D.K. Lee, T. Narasimhaswamy, R.P. Nanga, Cholesterol reduces pardaxin's

dynamics-a barrel-stave mechanism of membrane disruption investigated by solid-state NMR,

Biochim. Biophys. Acta, 1798 (2010) 223–227.

[49] C.J. Morton, M.A. Sani, M.W. Parker, F. Separovic, Cholesterol-Dependent Cytolysins:

Membrane and Protein Structural Requirements for Pore Formation, Chem. Rev., 119 (2019)

7721–7736.

[50] R.S. Cantor, Lipid composition and the lateral pressure profile in bilayers, Biophys. J., 76

(1999) 2625–2639.

[51] R.S. Cantor, Size distribution of barrel-stave aggregates of membrane peptides: influence of

the bilayer lateral pressure profile, Biophys. J., 82 (2002) 2520–2525.

[52] X. Feng, P. Barth, A topological and conformational stability alphabet for multipass membrane

proteins, Nat. Chem. Biol., 12 (2016) 167–173.

26

A Self-archived copy in

Kyoto University Research Information Repository

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

Table 1. Average lifetimes of monomers and dimers at 298 K

POPC

peptides

monomer lifetime

(ms)

POPC/cholesterol (7/3)

dimer lifetime

(ms)

monomer lifetime

(ms)

dimer lifetime

(ms)

----------------------------------------------------------------------------------------------------------------------------------------------1TM–1TM

not detected

not detected

562

261

1TM–2TM

545

99

159

381

2TM–2TM

1000

394

567

478

--------------------------------------------------------------------------------------------------------------------------------------------

27

A Self-archived copy in

Kyoto University Research Information Repository

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

Table 2. Thermodynamic parameters of the association of helices in POPC membranes at 298 K

Peptides

∆Ga (kJ mol–1)

∆Ha (kJ mol–1)

–T∆Sa (kJ mol–1)

-------------------------------------------- -------------------------------------------- -----------------------------------------------POPC

+chol

difference

POPC

+chol

difference

POPC

+chol

difference

--------------------------------------------------------------------------------------------------------------------------------------------------------------1TM–1TM

–13.2 ± 0.2* –22.6 ± 0.1** –9.4

–23.7 ± 0.4* –84.1 ± 1.7** –60.4

+10.4 ± 0.4* +61.4 ± 1.7** +51.0

1TM–2TM

–19.3 ± 1.2

–28.3 ± 2.6

–9.0

–35.4 ± 1.0

–59.3 ± 3.2

+16.1 ± 1.0

2TM–2TM

–21.6 ± 2.2

–24.2 ± 3.1

–2.6

–54.7 ± 6.6

–109.1 ± 20–54.4

–23.9

+33.4 ± 6.6

+31.0 ± 3.2

+84.0 ± 20

+14.9

+50.6

--------------------------------------------------------------------------------------------------------------------------------------------------------------*data from ensemble measurements (assuming ∆Cp(a) = –0.5 J K–1 mol–1)25

**data from ensemble measurements (assuming ∆Cp(a) = 1.5 J K–1 mol–1)23

28

A Self-archived copy in

Kyoto University Research Information Repository

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

Table 3. FTIR–PATR parameters for assessment of helix orientation at 298 K

POPC

Peptides

POPC/cholesterol(7/3)

α*

R*

(deg)

(deg)

------------------------------------------------------------------------------------------------------------------------------------------1TM–1TM

6.5 ± 0.6

1TM–2TM

4.8 ± 0.2

2TM–2TM

4.3 ± 0.1

~0

4.7 ± 0.2

17 ± 2

15 ± 3

3.6 ± 0.1

30 ± 1

22 ± 1

4.2 ± 0.1

23 ± 3

------------------------------------------------------------------------------------------------------------------------------------------* R and α indicate the dichroic ratio for the peptide amide I band and angle of helix orientation relative to the bilayer normal

calculated from R (Eq. 1), respectively.

29

A Self-archived copy in

Kyoto University Research Information Repository

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

FIGURE LEGENDS

Figure 1. Single-pair FRET (sp-FRET) measurements. (a) Schematic illustration of a

surface-attached vesicle for sp-FRET imaging by total internal reflection microscopy. (b)

Representative time-courses of fluorescence intensities for Cy3B (green) and Cy5 (Red)

under excitation of Cy3B for 1TM–1TM (left column), 1TM–2TM (middle column), and

2TM–2TM pairs (right column) in POPC (upper row) and POPC/cholesterol (lower row) at

25°C. The number of analyzed vesicles is denoted by n. Only Cy3B fluoresced for 1TM–

1TM in POPC, indicating no association between the helices. Under other conditions, the

emissions fluctuated with an anticorrelation, reflecting the association–dissociation dynamics

of the helices. (c)(d) Photobleaching of Cy3B (c) and Cy5 (d) detected as a stepwise

decrease in the fluorescence intensities (arrowheads). The vesicles that had incorporated one

Cy3B-helix and one Cy5-helix were selected for analysis.

Figure 2. Sp-FRET analysis for 1TM–2TM and 2TM–2TM associations in POPC vesicles.

(a) HaMMy fitting for sp-FRET trajectories assuming two-state transitions. Black and green

lines indicate measured apparent FRET efficiency (Eapp) and the most probable fitting,

respectively. The monomer and dimer states correspond to Eapp values of ~0.4 and ~0.8,

respectively. (b) Histograms of natural logarithm of rate constants for dimer formation (kon)

and dimer dissociation (koff). The average ± error values were obtained from Gaussian

fitting of the histogram. (c) Histograms of Eapp.

30

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 3. Sp-FRET analysis for 1TM–1TM, 1TM–2TM, and 2TM–2TM associations in

POPC/cholesterol (7/3) vesicles. (a) Histograms of the natural logarithm of rate constants for

dimer formation (kon) and dimer dissociation (koff). The average ± error values were

obtained from Gaussian fitting of the histogram. (b) Histograms of Eapp.

Figure 4. Temperature dependences of association free energy for (a) 1TM–2TM

associations and (b) 2TM–2TM associations. The temperature dependences were linearly

fitted to estimate the thermodynamic parameters by Eq. (7).

Figure 5. Amide region FTIR-PATR spectra for 1TM, 1TM/2TM(1/1) or 2TM incorporated

into membranes (peptides/lipids = 1/1000) at 25℃. The membrane films were hydrated with

D2O vapor. Red and black lines indicate raw spectra for the IR beam with its electric vector

parallel and perpendicular to the plane of incidence, respectively.

Figure 6. Effects of the membrane lateral pressure profile on the shape of the transmembrane

helix bundle. Because of the small headgroup of cholesterol, the headgroup and hydrocarbon

core regions have lower and higher lateral pressures, respectively, in cholesterol-containing

membranes (red arrows), compared with in pure POPC membranes. POPC membranes

stabilize helix bundles without helix tilt (parallel helix bundle). On the other hand,

cholesterol-containing membranes can stabilize helix bundles with significant helix tilt

(hourglass-shaped helix bundle).

31

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 7. Alterations in the helix macrodipole interactions upon the formation of the 1TM–

2TM bundle (a) and 2TM–2TM bundle (b). Upward and downward arrows in the circles

indicate C-terminus-up and C-terminus-down transmembrane topologies of the helices,

respectively. Antiparallel (A) and parallel (P or P’) interhelical contacts are shown as blue

and red lines, respectively. (a) 1TM–2TM association generates A + P interactions, which is

zero assuming symmetric packing of the helices in the bundle. (b) The repulsive interactions

in 2TM–2TM bundle are minimized in square packing (P’). The association generates 2A +

2P’ interactions, which have a negative ∆H value.

32

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 1

33

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 2

34

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 3

35

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 4

36

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 5

37

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 6

38

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure 7

39

A Self-archived copy in

Kyoto University Research Information Repository

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

Supporting information

Thermodynamic and kinetic stabilities of transmembrane helix bundles as

revealed by single-pair FRET analysis: Effects of the number of membranespanning segments and cholesterol

Yoshiaki Yano, Yuta Watanabe, and Katsumi Matsuzaki

NBD

Cy3B

Cy5

peptide

peptide

peptide

Figure S1. Structures of florescent probes used in this study.

A Self-archived copy in

Kyoto University Research Information Repository

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

Figure S2. HPLC chromatograms of transmembrane peptides. Purified fluorophore-labeled

peptides (I–VI) were analyzed with a PLRP-s analytical column (150×4.6 mm) at 50°C. The 1TM

peptides were eluted with a linear gradient of H2O/0.1% TFA and AcCN/0.1% TFA from 65 to 95%

(I), and from 10 to 90% (II, III). The 2TM peptides (IV, V, VI) were eluted with a linear gradient of

formic acid/H2O (2/3, v/v) and formic acid/2-propanol (4/1, v/v) from 50 to 80% (0–5min) and 80 to

95% (5–30 min).

...

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