• head_banner_01
  • head_banner_02

Crankshaft Position Sensor bakeng sa BMW, 12141247622 009163681 6PU009163681 17056 CS1330 EPS535 20936769 V20720475 75

Tlhaloso e Khutšoanyane:

Khouto ya Sehlahiswa:
YS-CKP1120
OEM:
E fumaneha
Mohlala:
E fumaneha
Tefo:
PayPal, Tse ling, VISA, MasterCard, Western Union, T/T
Sebaka sa Tšimoloho:
China
Matla a Phepelo:
50000 piece ka khoeli

Lintlha tsa Sehlahisoa

Li-tag tsa Sehlahisoa

Lintlha tse Potlakileng
  • TIISETSO
    1 SELEMO
Tefiso ea ho romella
Tefiso ea ho romella

BMW 3 COMPACT/COUPE (1990/09 – 1998/02)

BMW Z3 (1995/10 - 2003/01)


  • E fetileng:
  • E 'ngoe:

  • Ngola molaetsa wa hao mona mme o re romele wona