みぞ - 「ブリキノダンス」歌ってみた@溝

Edit History

Beat

  • 2020-06-18 21:52:21 +0900 (id: 3644747) Automatic music-understanding technologies
  • 2013-12-02 22:42:31 +0900 (id: 686937) Automatic music-understanding technologies Revert to this revision

Chord

  • 2013-12-02 22:42:19 +0900 (id: 688330) Automatic music-understanding technologies

Chorus

  • 2013-12-02 22:43:31 +0900 (id: 687887) Automatic music-understanding technologies

Melody

  • 2013-12-02 22:38:21 +0900 Automatic music-understanding technologies

Copy music map