Register      Login
Reproduction, Fertility and Development Reproduction, Fertility and Development Society
Vertebrate reproductive science and technology
RESEARCH ARTICLE

Growth arrest specific 1 (Gas1) and glial cell line-derived neurotrophic factor receptor α1 (Gfrα1), two mouse oocyte glycosylphosphatidylinositol-anchored proteins, are involved in fertilisation

M. Agopiantz A B C , L. Xandre-Rodriguez C , B. Jin C , G. Urbistondoy C , C. Ialy-Radio A B C , M. Chalbi A B C , J.-P. Wolf D , A. Ziyyat A B C and B. Lefèvre A B C E
+ Author Affiliations
- Author Affiliations

A Inserm, U1016, Institut Cochin, 24 rue du Faubourg Saint-Jacques, 75014, Paris, France.

B CNRS, UMR8104, 24 rue du Faubourg Saint-Jacques, 75014, Paris, France.

C Université Paris Descartes, Sorbonne Paris Cité, 24 rue du Faubourg Saint-Jacques, 75014, Paris, France.

D Service d’Histologie Embryologie Biologie de la Reproduction – CECOS, Hôpital Cochin, AP-HP, F75014 Paris, France.

E Corresponding author. Email: brigitte.lefevre@parisdescartes.fr

Reproduction, Fertility and Development 29(4) 824-837 https://doi.org/10.1071/RD15367
Submitted: 8 September 2015  Accepted: 10 December 2015   Published: 24 February 2016

Abstract

Recently, Juno, the oocyte receptor for Izumo1, a male immunoglobulin, was discovered. Juno is an essential glycosylphosphatidylinositol (GIP)-anchored protein. This result did not exclude the participation of other GIP-anchored proteins in this process. After bibliographic and database searches we selected five GIP-anchored proteins (Cpm, Ephrin-A4, Gas1, Gfra1 and Rgmb) as potential oocyte candidates participating in fertilisation. Western blot and immunofluorescence analyses showed that only three were present on the mouse ovulated oocyte membrane and, of these, only two were clearly involved in the fertilisation process, namely growth arrest specific 1 (Gas1) and glial cell line-derived neurotrophic factor receptor α1 (Gfrα1). This was demonstrated by evaluating oocyte fertilisability after treatment of oocytes with antibodies against the selected proteins, with their respective short interference RNA or both. Gfrα1 and Gas1 seem to be neither redundant nor synergistic. In conclusion, oocyte Gas1 and Gfrα1 are both clearly involved in fertilisation.

Additional keywords: short interference RNA.


References

Adeyo, O., Allan, B.B., Barnes, R.H., Goulbourne, C.N., Tatar, A., Tu, Y., Young, L.C., Weinstein, M.M., Tontonoz, P., Fong, L.G., Beigneux, A.P., and Young, S.G. (2014). Palmoplantar keratoderma along with neuromuscular and metabolic phenotypes in Slurp1-deficient mice. J. Invest. Dermatol. 134, 1589–1598.
Palmoplantar keratoderma along with neuromuscular and metabolic phenotypes in Slurp1-deficient mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXhsl2msLk%3D&md5=e3da2727849b7fc385f2393b60571b2aCAS | 24499735PubMed |

Airaksinen, M. S., Titievsky, A., and Saarma, M. (1999). GDNF family neurotrophic factor signaling: four masters, one servant? Mol. Cell. Neurosci. 13, 313–325.
GDNF family neurotrophic factor signaling: four masters, one servant?Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXjsVGjsb4%3D&md5=8c1cd0f6e09840b5cb07f89aea917529CAS | 10356294PubMed |

Alfieri, J. A., Martin, A. D., Takeda, J., Kondoh, G., Myles, D. G., and Primakoff, P. (2003). Infertility in female mice with an oocyte-specific knockout of GPI-anchored proteins. J. Cell Sci. 116, 2149–2155.
Infertility in female mice with an oocyte-specific knockout of GPI-anchored proteins.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXkslCltLc%3D&md5=1b6b268263061d320a898eab255bb349CAS | 12692150PubMed |

Amanai, M., Shoji, S., Yoshida, N., Brahmajosyula, M., and Perry, A. C. (2006). Injection of mammalian metaphase II oocytes with short interfering RNAs to dissect meiotic and early mitotic events. Biol. Reprod. 75, 891–898.
Injection of mammalian metaphase II oocytes with short interfering RNAs to dissect meiotic and early mitotic events.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28Xht1yjsbfF&md5=0247c953a48befe070b4cf1a4a995942CAS | 16943363PubMed |

Arnauld, S., Fidaleo, M., Clemencet, M. C., Chevillard, G., Athias, A., Gresti, J., Wanders, R. J., Latruffe, N., Nicolas-Frances, V., and Mandard, S. (2009). Modulation of the hepatic fatty acid pool in peroxisomal 3-ketoacyl-CoA thiolase B-null mice exposed to the selective PPARalpha agonist Wy14,643. Biochimie 91, 1376–1386.
Modulation of the hepatic fatty acid pool in peroxisomal 3-ketoacyl-CoA thiolase B-null mice exposed to the selective PPARalpha agonist Wy14,643.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhsVWmurfN&md5=1c17a46536072599c114e136676fe6dbCAS | 19772884PubMed |

Bates, J. M., Raffi, H. M., Prasadan, K., Mascarenhas, R., Laszik, Z., Maeda, N., Hultgren, S. J., and Kumar, S. (2004). Tamm–Horsfall protein knockout mice are more prone to urinary tract infection: rapid communication. Kidney Int. 65, 791–797.
Tamm–Horsfall protein knockout mice are more prone to urinary tract infection: rapid communication.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXitleks7k%3D&md5=7695dd68673f603ae1c99bad160a6e8eCAS | 14871399PubMed |

Beigneux, A. P., Davies, B. S., Gin, P., Weinstein, M. M., Farber, E., Qiao, X., Peale, F., Bunting, S., Walzem, R. L., Wong, J. S., Blaner, W. S., Ding, Z. M., Melford, K., Wongsiriroj, N., Shu, X., de Sauvage, F., Ryan, R. O., Fong, L. G., Bensadoun, A., and Young, S. G. (2007). Glycosylphosphatidylinositol-anchored high-density lipoprotein-binding protein 1 plays a critical role in the lipolytic processing of chylomicrons. Cell Metab. 5, 279–291.
Glycosylphosphatidylinositol-anchored high-density lipoprotein-binding protein 1 plays a critical role in the lipolytic processing of chylomicrons.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXkslGmsLc%3D&md5=047e3a0ebee915ec84319ac3a1413006CAS | 17403372PubMed |

Belvindrah, R., Rougon, G., and Chazal, G. (2002). Increased neurogenesis in adult mCD24-deficient mice. J. Neurosci. 22, 3594–3607.
| 1:CAS:528:DC%2BD38Xjs1Sgs7Y%3D&md5=f1d86e6b839e953c51cb0e83ce0b64e6CAS | 11978835PubMed |

Berglund, E. O., Murai, K. K., Fredette, B., Sekerkova, G., Marturano, B., Weber, L., Mugnaini, E., and Ranscht, B. (1999). Ataxia and abnormal cerebellar microorganization in mice with ablated contactin gene expression. Neuron 24, 739–750.
Ataxia and abnormal cerebellar microorganization in mice with ablated contactin gene expression.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXnvFyjsrs%3D&md5=282973b0752d74d102ae2600ebc1eae0CAS | 10595523PubMed |

Bianchi, E., Doe, B., Goulding, D., and Wright, G. J. (2014). Juno is the egg Izumo receptor and is essential for mammalian fertilization. Nature 508, 483–487.
Juno is the egg Izumo receptor and is essential for mammalian fertilization.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC2cXmslaltb4%3D&md5=9d80524db8c884dc5956ecde0f15407eCAS | 24739963PubMed |

Bielke, W., Ke, G., Feng, Z., Buhrer, S., Saurer, S., and Friis, R. R. (1997). Apoptosis in the rat mammary gland and ventral prostate: detection of cell death-associated genes using a coincident-expression cloning approach. Cell Death Differ. 4, 114–124.
Apoptosis in the rat mammary gland and ventral prostate: detection of cell death-associated genes using a coincident-expression cloning approach.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK2sXotVehs7w%3D&md5=c526b68d034508cbb84eb378d749ee9cCAS | 16465217PubMed |

Boerke, A., van der Lit, J., Lolicato, F., Stout, T. A., Helms, J. B., and Gadella, B. M. (2014). Removal of GPI-anchored membrane proteins causes clustering of lipid microdomains in the apical head area of porcine sperm. Theriogenology 81, 613–624.
Removal of GPI-anchored membrane proteins causes clustering of lipid microdomains in the apical head area of porcine sperm.Crossref | GoogleScholarGoogle Scholar | 24377861PubMed |

Buschiazzo, J., Ialy-Radio, C., Auer, J., Wolf, J. P., Serres, C., Lefevre, B., and Ziyyat, A. (2013). Cholesterol depletion disorganizes oocyte membrane rafts altering mouse fertilization. PLoS One 8, e62919.
Cholesterol depletion disorganizes oocyte membrane rafts altering mouse fertilization.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXntlClsro%3D&md5=8cf22ecfecb7105d1d10667128bbf195CAS | 23638166PubMed |

Cabrera, J. R., Sanchez-Pulido, L., Rojas, A. M., Valencia, A., Manes, S., Naranjo, J. R., and Mellstrom, B. (2006). Gas1 is related to the glial cell-derived neurotrophic factor family receptors alpha and regulates Ret signaling. J. Biol. Chem. 281, 14 330–14 339.
Gas1 is related to the glial cell-derived neurotrophic factor family receptors alpha and regulates Ret signaling.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28XksFeitro%3D&md5=a282cf325c2a1b9b756abc6fd2511c7bCAS |

Cano-Gauci, D. F., Song, H. H., Yang, H., McKerlie, C., Choo, B., Shi, W., Pullano, R., Piscione, T. D., Grisaru, S., Soon, S., Sedlackova, L., Tanswell, A. K., Mak, T. W., Yeger, H., Lockwood, G. A., Rosenblum, N. D., and Filmus, J. (1999). Glypican-3-deficient mice exhibit developmental overgrowth and some of the abnormalities typical of Simpson–Golabi–Behmel syndrome. J. Cell Biol. 146, 255–264.
Glypican-3-deficient mice exhibit developmental overgrowth and some of the abnormalities typical of Simpson–Golabi–Behmel syndrome.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXksFCjsL8%3D&md5=e2573c93fdb03eedf57dded60af6f7a1CAS | 10402475PubMed |

Carmeliet, P., Kieckens, L., Schoonjans, L., Ream, B., van Nuffelen, A., Prendergast, G., Cole, M., Bronson, R., Collen, D., and Mulligan, R. C. (1993). Plasminogen activator inhibitor-1 gene-deficient mice. I. Generation by homologous recombination and characterization. J. Clin. Invest. 92, 2746–2755.
Plasminogen activator inhibitor-1 gene-deficient mice. I. Generation by homologous recombination and characterization.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK2cXltV2iug%3D%3D&md5=5065d2d1a1b751a36e022b67bef03637CAS | 8254028PubMed |

Catania, E. H., Pimenta, A., and Levitt, P. (2008). Genetic deletion of Lsamp causes exaggerated behavioral activation in novel environments. Behav. Brain Res. 188, 380–390.
| 1:CAS:528:DC%2BD1cXhslOju7w%3D&md5=97695559c80b4be263c09bd720a905a6CAS | 18199495PubMed |

Cheng, H., Govindan, J. A., and Greenstein, D. (2008). Regulated trafficking of the MSP/Eph receptor during oocyte meiotic maturation in C. elegans. Curr. Biol. 18, 705–714.
Regulated trafficking of the MSP/Eph receptor during oocyte meiotic maturation in C. elegans.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXmtVaktro%3D&md5=76dbed17963493c2c1d349002207b1d6CAS | 18472420PubMed |

Coonrod, S. A., Naaby-Hansen, S., Shetty, J., Shibahara, H., Chen, M., White, J. M., and Herr, J. C. (1999). Treatment of mouse oocytes with PI-PLC releases 70-kDa (pI 5) and 35- to 45-kDa (pI 5.5) protein clusters from the egg surface and inhibits sperm–oolemma binding and fusion. Dev. Biol. 207, 334–349.
Treatment of mouse oocytes with PI-PLC releases 70-kDa (pI 5) and 35- to 45-kDa (pI 5.5) protein clusters from the egg surface and inhibits sperm–oolemma binding and fusion.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXhsFajtLo%3D&md5=1b512e9dd03f2dc7507970cd0eea8102CAS | 10068467PubMed |

Corrigan, C., Subramanian, R., and Miller, M. A. (2005). Eph and NMDA receptors control Ca2+/calmodulin-dependent protein kinase II activation during C. elegans oocyte meiotic maturation. Development 132, 5225–5237.
Eph and NMDA receptors control Ca2+/calmodulin-dependent protein kinase II activation during C. elegans oocyte meiotic maturation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28Xktlehsw%3D%3D&md5=bbcae564132cb100064d1e6ddb1cff64CAS | 16267094PubMed |

Cremer, H., Lange, R., Christoph, A., Plomann, M., Vopper, G., Roes, J., Brown, R., Baldwin, S., Kraemer, P., Scheff, S., Barthels, D., Rajewsky, K., and Wille, W. (1994). Inactivation of the N-CAM gene in mice results in size reduction of the olfactory bulb and deficits in spatial learning. Nature 367, 455–459.
Inactivation of the N-CAM gene in mice results in size reduction of the olfactory bulb and deficits in spatial learning.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK2cXhsF2qs70%3D&md5=67bc8b4e0d0176037b8f0a9360148659CAS | 8107803PubMed |

Cutforth, T., Moring, L., Mendelsohn, M., Nemes, A., Shah, N. M., Kim, M. M., Frisen, J., and Axel, R. (2003). Axonal ephrin-As and odorant receptors: coordinate determination of the olfactory sensory map. Cell 114, 311–322.
Axonal ephrin-As and odorant receptors: coordinate determination of the olfactory sensory map.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXmsFWqtrk%3D&md5=83d0710f0871ec53f64445aed2bd2c02CAS | 12914696PubMed |

Daude, N., Wohlgemuth, S., Brown, R., Pitstick, R., Gapeshina, H., Yang, J., Carlson, G. A., and Westaway, D. (2012). Knockout of the prion protein (PrP)-like Sprn gene does not produce embryonic lethality in combination with PrP(C)-deficiency. Proc. Natl Acad. Sci. USA 109, 9035–9040.
Knockout of the prion protein (PrP)-like Sprn gene does not produce embryonic lethality in combination with PrP(C)-deficiency.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XovF2gtbc%3D&md5=375066204f928686dc398b9f63ecc977CAS | 22619325PubMed |

DeChiara, T. M., Vejsada, R., Poueymirou, W. T., Acheson, A., Suri, C., Conover, J. C., Friedman, B., McClain, J., Pan, L., Stahl, N., Ip, N. Y., and Yancopoulos, G. D. (1995). Mice lacking the CNTF receptor, unlike mice lacking CNTF, exhibit profound motor neuron deficits at birth. Cell 83, 313–322.
Mice lacking the CNTF receptor, unlike mice lacking CNTF, exhibit profound motor neuron deficits at birth.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK2MXovVGrurg%3D&md5=ca88acce9e0ba01976320d05e5c90d42CAS | 7585948PubMed |

Deiteren, K., Hendriks, D., Scharpe, S., and Lambeir, A. M. (2009). Carboxypeptidase M: multiple alliances and unknown partners. Clin. Chim. Acta 399, 24–39.
Carboxypeptidase M: multiple alliances and unknown partners.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXhsVOisrfP&md5=5560546eab9133cb8f53a8e59f7c8f5eCAS | 18957287PubMed |

Del Sal, G., Ruaro, M. E., Philipson, L., and Schneider, C. (1992). The growth arrest-specific gene, gas1, is involved in growth suppression. Cell 70, 595–607.
The growth arrest-specific gene, gas1, is involved in growth suppression.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK3sXkt1Wkt70%3D&md5=e50b432103577e22f921dbf1d0626209CAS | 1505026PubMed |

Dickendesher, T. L., Baldwin, K. T., Mironova, Y. A., Koriyama, Y., Raiker, S. J., Askew, K. L., Wood, A., Geoffroy, C. G., Zheng, B., Liepmann, C. D., Katagiri, Y., Benowitz, L. I., Geller, H. M., and Giger, R. J. (2012). NgR1 and NgR3 are receptors for chondroitin sulfate proteoglycans. Nat. Neurosci. 15, 703–712.
NgR1 and NgR3 are receptors for chondroitin sulfate proteoglycans.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38Xjs1WitLc%3D&md5=af4b38a6a4e73fbe1e051367f6433e1cCAS | 22406547PubMed |

Ding, J., Yang, L., Yan, Y. T., Chen, A., Desai, N., Wynshaw-Boris, A., and Shen, M. M. (1998). Cripto is required for correct orientation of the anterior–posterior axis in the mouse embryo. Nature 395, 702–707.
Cripto is required for correct orientation of the anterior–posterior axis in the mouse embryo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1cXms1Ojsbo%3D&md5=f8d4c21ebe8fbf3e30f610d234759c45CAS | 9790191PubMed |

Ding, J., Swain, J. E., and Smith, G. D. (2011). Aurora kinase-A regulates microtubule organizing center (MTOC) localization, chromosome dynamics, and histone-H3 phosphorylation in mouse oocytes. Mol. Reprod. Dev. 78, 80–90.
Aurora kinase-A regulates microtubule organizing center (MTOC) localization, chromosome dynamics, and histone-H3 phosphorylation in mouse oocytes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXitVyltLw%3D&md5=ff6469d0d1491e5b498b194e976694d3CAS | 21274965PubMed |

Dole, G., Nilsson, E. E., and Skinner, M. K. (2008). Glial-derived neurotrophic factor promotes ovarian primordial follicle development and cell–cell interactions during folliculogenesis. Reproduction 135, 671–682.
Glial-derived neurotrophic factor promotes ovarian primordial follicle development and cell–cell interactions during folliculogenesis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXls1Wqur0%3D&md5=a9e98da50b678a6515247ec9ec2991cdCAS | 18304989PubMed |

Euteneuer, S., Yang, K. H., Chavez, E., Leichtle, A., Loers, G., Olshansky, A., Pak, K., Schachner, M., and Ryan, A. F. (2013). Glial cell line-derived neurotrophic factor (GDNF) induces neuritogenesis in the cochlear spiral ganglion via neural cell adhesion molecule (NCAM). Mol. Cell. Neurosci. 54, 30–43.
Glial cell line-derived neurotrophic factor (GDNF) induces neuritogenesis in the cochlear spiral ganglion via neural cell adhesion molecule (NCAM).Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3sXlt1Cgtrk%3D&md5=482247b6afb103aa883a6ebf66366b57CAS | 23262364PubMed |

Farhi, J., Ao, A., Fisch, B., Zhang, X. Y., Garor, R., and Abir, R. (2010). Glial cell line-derived neurotrophic factor (GDNF) and its receptors in human ovaries from fetuses, girls, and women. Fertil. Steril. 93, 2565–2571.
Glial cell line-derived neurotrophic factor (GDNF) and its receptors in human ovaries from fetuses, girls, and women.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3cXot1aruro%3D&md5=a4a66278377b01213e0f12a0da5b5031CAS | 19896648PubMed |

Feng, G., Laskowski, M. B., Feldheim, D. A., Wang, H., Lewis, R., Frisen, J., Flanagan, J. G., and Sanes, J. R. (2000). Roles for ephrins in positionally selective synaptogenesis between motor neurons and muscle fibers. Neuron 25, 295–306.
Roles for ephrins in positionally selective synaptogenesis between motor neurons and muscle fibers.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXhslOktr0%3D&md5=6e10c8e53f67445e8965df3351f726e6CAS | 10719886PubMed |

Fischer, M., Rulicke, T., Raeber, A., Sailer, A., Moser, M., Oesch, B., Brandner, S., Aguzzi, A., and Weissmann, C. (1996). Prion protein (PrP) with amino-proximal deletions restoring susceptibility of PrP knockout mice to scrapie. EMBO J. 15, 1255–1264.
| 1:CAS:528:DyaK28XitFWjtL4%3D&md5=5813911d01011eb281e2ede675b7a227CAS | 8635458PubMed |

Frieden, L. A., Townsend, T. A., Vaught, D. B., Delaughter, D. M., Hwang, Y., Barnett, J. V., and Chen, J. (2010). Regulation of heart valve morphogenesis by Eph receptor ligand, ephrin-A1. Dev. Dyn. 239, 3226–3234.
Regulation of heart valve morphogenesis by Eph receptor ligand, ephrin-A1.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXitlartQ%3D%3D&md5=3b1cc002a14174575d3076b77ce87a8fCAS | 20960543PubMed |

Frisén, J., Yates, P. A., McLaughlin, T., Friedman, G. C., O’Leary, D. D., and Barbacid, M. (1998). Ephrin-A5 (AL-1/RAGS) is essential for proper retinal axon guidance and topographic mapping in the mammalian visual system. Neuron 20, 235–243.
Ephrin-A5 (AL-1/RAGS) is essential for proper retinal axon guidance and topographic mapping in the mammalian visual system.Crossref | GoogleScholarGoogle Scholar | 9491985PubMed |

Fujihara, Y., Okabe, M., and Ikawa, M. (2014). GPI-anchored protein complex, LY6K/TEX101, is required for sperm migration into the oviduct and male fertility in mice. Biol. Reprod. 90, 60.
GPI-anchored protein complex, LY6K/TEX101, is required for sperm migration into the oviduct and male fertility in mice.Crossref | GoogleScholarGoogle Scholar | 24501175PubMed |

Fujii, H., Tatsumi, K., Kosaka, K., Yoshioka, S., Fujiwara, H., and Fujii, S. (2006). Eph–ephrin A system regulates murine blastocyst attachment and spreading. Dev. Dyn. 235, 3250–3258.
Eph–ephrin A system regulates murine blastocyst attachment and spreading.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXmsVWjtA%3D%3D&md5=4edd4e21e2bee03566af6100279614dbCAS | 17039519PubMed |

Fukamauchi, F., Aihara, O., Wang, Y. J., Akasaka, K., Takeda, Y., Horie, M., Kawano, H., Sudo, K., Asano, M., Watanabe, K., and Iwakura, Y. (2001). TAG-1-deficient mice have marked elevation of adenosine A1 receptors in the hippocampus. Biochem. Biophys. Res. Commun. 281, 220–226.
TAG-1-deficient mice have marked elevation of adenosine A1 receptors in the hippocampus.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXhtFemt74%3D&md5=5093d27d64a116f58ab670b061182b5bCAS | 11178983PubMed |

Ibáñez, C. F. (2010). Beyond the cell surface: new mechanisms of receptor function. Biochem. Biophys. Res. Commun. 396, 24–27.
Beyond the cell surface: new mechanisms of receptor function.Crossref | GoogleScholarGoogle Scholar | 20494105PubMed |

Inoue, N., Ikawa, M., Isotani, A., and Okabe, M. (2005). The immunoglobulin superfamily protein Izumo is required for sperm to fuse with eggs. Nature 434, 234–238.
The immunoglobulin superfamily protein Izumo is required for sperm to fuse with eggs.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXitV2hsrs%3D&md5=00fabc0f5fade434f01d0c224671f5eeCAS | 15759005PubMed |

Jégou, A., Ziyyat, A., Barraud-Lange, V., Perez, E., Wolf, J. P., Pincet, F., and Gourier, C. (2011). CD9 tetraspanin generates fusion competent sites on the egg membrane for mammalian fertilization. Proc. Natl Acad. Sci. USA 108, 10 946–10 951.
CD9 tetraspanin generates fusion competent sites on the egg membrane for mammalian fertilization.Crossref | GoogleScholarGoogle Scholar |

Jia, Z., Zhao, R., Tian, Y., Huang, Z., Tian, Z., Shen, Z., Wang, Q., Wang, J., Fu, X., and Wu, Y. (2009). A novel splice variant of FR4 predominantly expressed in CD4+CD25+ regulatory T cells. Immunol. Invest. 38, 718–729.
A novel splice variant of FR4 predominantly expressed in CD4+CD25+ regulatory T cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhsFSqsbvO&md5=878105aa7c5d697994084b485f9b65e4CAS | 19860584PubMed |

Jiang, Z., Xu, Y., and Cai, S. (2011). Down-regulated GAS1 expression correlates with recurrence in stage II and III colorectal cancer. Hum. Pathol. 42, 361–368.
Down-regulated GAS1 expression correlates with recurrence in stage II and III colorectal cancer.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXhvFWgtLg%3D&md5=300effdb9127db93e925ee6e1c209da3CAS | 21111449PubMed |

González-Cabrero, J., Wise, C. J., Latchman, Y., Freeman, G. J., Sharpe, A. H., and Reiser, H. (1999). CD48-deficient mice have a pronounced defect in CD4(+) T cell activation. Proc. Natl Acad. Sci. USA 96, 1019–1023.
CD48-deficient mice have a pronounced defect in CD4(+) T cell activation.Crossref | GoogleScholarGoogle Scholar | 9927686PubMed |

Grigorieva, A., Griffiths, G. S., Zhang, H., Laverty, G., Shao, M., Taylor, L., and Martin-DeLeon, P. A. (2007). Expression of SPAM1 (PH-20) in the murine kidney is not accompanied by hyaluronidase activity: evidence for potential roles in fluid and water reabsorption. Kidney Blood Press. Res. 30, 145–155.
Expression of SPAM1 (PH-20) in the murine kidney is not accompanied by hyaluronidase activity: evidence for potential roles in fluid and water reabsorption.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXmvFSmsrs%3D&md5=7898f305b30e4b7a2e4accc007caad73CAS | 17446714PubMed |

Guerra, N., Tan, Y. X., Joncker, N. T., Choy, A., Gallardo, F., Xiong, N., Knoblaugh, S., Cado, D., Greenberg, N. M., and Raulet, D. H. (2008). NKG2D-deficient mice are defective in tumor surveillance in models of spontaneous malignancy. Immunity 28, 571–580.
NKG2D-deficient mice are defective in tumor surveillance in models of spontaneous malignancy.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXkvV2itL8%3D&md5=45096e5ce16ed14f4d6588d6b1c1042cCAS | 18394936PubMed |

Hazenbos, W. L., Gessner, J. E., Hofhuis, F. M., Kuipers, H., Meyer, D., Heijnen, I. A., Schmidt, R. E., Sandor, M., Capel, P. J., Daeron, M., van de Winkel, J. G., and Verbeek, J. S. (1996). Impaired IgG-dependent anaphylaxis and Arthus reaction in Fc gamma RIII (CD16) deficient mice. Immunity 5, 181–188.
Impaired IgG-dependent anaphylaxis and Arthus reaction in Fc gamma RIII (CD16) deficient mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK28XltlKjsbk%3D&md5=b5fed78d8cecb896aabd5ee8b578bea8CAS | 8769481PubMed |

Haziot, A., Ferrero, E., Lin, X. Y., Stewart, C. L., and Goyert, S. M. (1995). CD14-deficient mice are exquisitely insensitive to the effects of LPS. Prog. Clin. Biol. Res. 392, 349–351.
| 1:CAS:528:DyaK28XmsFOlsbs%3D&md5=e72b2b8a562f0e828791e025439a67e6CAS | 8524940PubMed |

Hebbard, L. W., Garlatti, M., Young, L. J., Cardiff, R. D., Oshima, R. G., and Ranscht, B. (2008). T-Cadherin supports angiogenesis and adiponectin association with the vasculature in a mouse mammary tumor model. Cancer Res. 68, 1407–1416.
T-Cadherin supports angiogenesis and adiponectin association with the vasculature in a mouse mammary tumor model.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXislagurg%3D&md5=4419cafa2af50435fa864f9535969f8bCAS | 18316604PubMed |

Hendrich, B., Guy, J., Ramsahoye, B., Wilson, V. A., and Bird, A. (2001). Closely related proteins MBD2 and MBD3 play distinctive but interacting roles in mouse development. Genes Dev. 15, 710–723.
Closely related proteins MBD2 and MBD3 play distinctive but interacting roles in mouse development.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXisFSku74%3D&md5=3eff469c60b73b4d930e4704ab56eb8cCAS | 11274056PubMed |

Honma, Y., Araki, T., Gianino, S., Bruce, A., Heuckeroth, R., Johnson, E., and Milbrandt, J. (2002). Artemin is a vascular-derived neurotropic factor for developing sympathetic neurons. Neuron 35, 267–282.
Artemin is a vascular-derived neurotropic factor for developing sympathetic neurons.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38XlvVCmu70%3D&md5=7f9b310d924f5f0d9287604000aefb2eCAS | 12160745PubMed |

Horn, K. H., Esposito, E. R., Greene, R. M., and Pisano, M. M. (2008). The effect of cigarette smoke exposure on developing folate binding protein-2 null mice. Reprod. Toxicol. 26, 203–209.
The effect of cigarette smoke exposure on developing folate binding protein-2 null mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXhtlyhsL%2FO&md5=b2ce18653637c176ca6502226a2f430aCAS | 18992323PubMed |

Itoh, M., Ishihara, K., Hiroi, T., Lee, B. O., Maeda, H., Iijima, H., Yanagita, M., Kiyono, H., and Hirano, T. (1998). Deletion of bone marrow stromal cell antigen-1 (CD157) gene impaired systemic thymus independent-2 antigen-induced IgG3 and mucosal TD antigen-elicited IgA responses. J. Immunol. 161, 3974–3983.
| 1:CAS:528:DyaK1cXmsFyrt7c%3D&md5=5bd758319e5e48ae9e78194386518e4dCAS | 9780166PubMed |

Jen, Y. H., Musacchio, M., and Lander, A. D. (2009). Glypican-1 controls brain size through regulation of fibroblast growth factor signaling in early neurogenesis. Neural Dev. 4, 33.
Glypican-1 controls brain size through regulation of fibroblast growth factor signaling in early neurogenesis.Crossref | GoogleScholarGoogle Scholar | 19732411PubMed |

Ji, B., Case, L. C., Liu, K., Shao, Z., Lee, X., Yang, Z., Wang, J., Tian, T., Shulga-Morskaya, S., Scott, M., He, Z., Relton, J. K., and Mi, S. (2008). Assessment of functional recovery and axonal sprouting in oligodendrocyte-myelin glycoprotein (OMgp) null mice after spinal cord injury. Mol. Cell. Neurosci. 39, 258–267.
Assessment of functional recovery and axonal sprouting in oligodendrocyte-myelin glycoprotein (OMgp) null mice after spinal cord injury.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXhtFCnsrzP&md5=c2f59a8d56ce7b076e0e420885d44f33CAS | 18692574PubMed |

Kaji, K., Oda, S., Shikano, T., Ohnuki, T., Uematsu, Y., Sakagami, J., Tada, N., Miyazaki, S., and Kudo, A. (2000). The gamete fusion process is defective in eggs of Cd9-deficient mice. Nat. Genet. 24, 279–282.
The gamete fusion process is defective in eggs of Cd9-deficient mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXhvFaqt7s%3D&md5=e88ae267ed96022958398be84ef81a06CAS | 10700183PubMed |

Kawamura, K., Ye, Y., Kawamura, N., Jing, L., Groenen, P., Gelpke, M. S., Rauch, R., Hsueh, A. J., and Tanaka, T. (2008). Completion of meiosis I of preovulatory oocytes and facilitation of preimplantation embryo development by glial cell line-derived neurotrophic factor. Dev. Biol. 315, 189–202.
Completion of meiosis I of preovulatory oocytes and facilitation of preimplantation embryo development by glial cell line-derived neurotrophic factor.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXitlKksrc%3D&md5=e15208c8795c4c8419a703e709e78045CAS | 18234170PubMed |

Keller, S. R., Davis, A. C., and Clairmont, K. B. (2002). Mice deficient in the insulin-regulated membrane aminopeptidase show substantial decreases in glucose transporter GLUT4 levels but maintain normal glucose homeostasis. J. Biol. Chem. 277, 17 677–17 686.
Mice deficient in the insulin-regulated membrane aminopeptidase show substantial decreases in glucose transporter GLUT4 levels but maintain normal glucose homeostasis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38XktVCnsrs%3D&md5=b7d9c4c5a4f2e819a0d76904e625a388CAS |

Killeen, N., Stuart, S. G., and Littman, D. R. (1992). Development and function of T cells in mice with a disrupted CD2 gene. EMBO J. 11, 4329–4336.
| 1:CAS:528:DyaK3sXitV2jsQ%3D%3D&md5=5997e6ce880943ac5c86bb30a2414cffCAS | 1358605PubMed |

Kim, J. E., Liu, B. P., Park, J. H., and Strittmatter, S. M. (2004). Nogo-66 receptor prevents raphespinal and rubrospinal axon regeneration and limits functional recovery from spinal cord injury. Neuron 44, 439–451.
Nogo-66 receptor prevents raphespinal and rubrospinal axon regeneration and limits functional recovery from spinal cord injury.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXpvVChs7s%3D&md5=ba2fc9dfe628deb58ffea3084b4980ffCAS | 15504325PubMed |

Kimura, M., Kim, E., Kang, W., Yamashita, M., Saigo, M., Yamazaki, T., Nakanishi, T., Kashiwabara, S., and Baba, T. (2009). Functional roles of mouse sperm hyaluronidases, HYAL5 and SPAM1, in fertilization. Biol. Reprod. 81, 939–947.
Functional roles of mouse sperm hyaluronidases, HYAL5 and SPAM1, in fertilization.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhtlWrtrzI&md5=04532f4efbbad2b0f992f5471f3c6df8CAS | 19605784PubMed |

Koszalka, P., Ozuyaman, B., Huo, Y., Zernecke, A., Flogel, U., Braun, N., Buchheiser, A., Decking, U. K., Smith, M. L., Sevigny, J., Gear, A., Weber, A. A., Molojavyi, A., Ding, Z., Weber, C., Ley, K., Zimmermann, H., Godecke, A., and Schrader, J. (2004). Targeted disruption of cd73/ecto-5′-nucleotidase alters thromboregulation and augments vascular inflammatory response. Circ. Res. 95, 814–821.
Targeted disruption of cd73/ecto-5′-nucleotidase alters thromboregulation and augments vascular inflammatory response.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXot1aisbY%3D&md5=a8a22912f607163b0c42577ff8970254CAS | 15358667PubMed |

Koyama, K., Hasegawa, A., and Komori, S. (2009). Functional aspects of CD52 in reproduction. J. Reprod. Immunol. 83, 56–59.
Functional aspects of CD52 in reproduction.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXhsVyntrrL&md5=f45afd2b0e06c052be9d821c1c42347aCAS | 19875176PubMed |

Le Naour, F., Rubinstein, E., Jasmin, C., Prenant, M., and Boucheix, C. (2000). Severely reduced female fertility in CD9-deficient mice. Science 287, 319–321.
Severely reduced female fertility in CD9-deficient mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXlvVOntA%3D%3D&md5=641dded137d94de89b0e8c63d2733225CAS | 10634790PubMed |

Lee, A. W., Hengstler, H., Schwald, K., Berriel-Diaz, M., Loreth, D., Kirsch, M., Kretz, O., Haas, C. A., de Angelis, M. H., Herzig, S., Brummendorf, T., Klingenspor, M., Rathjen, F. G., Rozman, J., Nicholson, G., Cox, R. D., and Schafer, M. K. (2012). Functional inactivation of the genome-wide association study obesity gene neuronal growth regulator 1 in mice causes a body mass phenotype. PLoS One 7, e41537.
Functional inactivation of the genome-wide association study obesity gene neuronal growth regulator 1 in mice causes a body mass phenotype.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC38XhtFSkur7E&md5=3fc478dd7394de0519d133914ee37ddbCAS | 22844493PubMed |

Lefèvre, B., Wolf, J. P., and Ziyyat, A. (2010). Sperm–egg interaction: is there a link between tetraspanin(s) and GPI-anchored protein(s)? BioEssays 32, 143–152.
Sperm–egg interaction: is there a link between tetraspanin(s) and GPI-anchored protein(s)?Crossref | GoogleScholarGoogle Scholar | 20091756PubMed |

Legan, P. K., Lukashkina, V. A., Goodyear, R. J., Kossi, M., Russell, I. J., and Richardson, G. P. (2000). A targeted deletion in alpha-tectorin reveals that the tectorial membrane is required for the gain and timing of cochlear feedback. Neuron 28, 273–285.
A targeted deletion in alpha-tectorin reveals that the tectorial membrane is required for the gain and timing of cochlear feedback.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXnvVGltb4%3D&md5=877336f486256a1a85eca7569dc678e6CAS | 11087000PubMed |

Leppilampi, M., Karttunen, T. J., Kivela, J., Gut, M. O., Pastorekova, S., Pastorek, J., and Parkkila, S. (2005). Gastric pit cell hyperplasia and glandular atrophy in carbonic anhydrase IX knockout mice: studies on two strains C57/BL6 and BALB/c. Transgenic Res. 14, 655–663.
Gastric pit cell hyperplasia and glandular atrophy in carbonic anhydrase IX knockout mice: studies on two strains C57/BL6 and BALB/c.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXhtFemu7nL&md5=6990f8e5d895846970b778f8efff3fa3CAS | 16245156PubMed |

Lindfors, P. H., Lindahl, M., Rossi, J., Saarma, M., and Airaksinen, M. S. (2006). Ablation of persephin receptor glial cell line-derived neurotrophic factor family receptor alpha4 impairs thyroid calcitonin production in young mice. Endocrinology 147, 2237–2244.
Ablation of persephin receptor glial cell line-derived neurotrophic factor family receptor alpha4 impairs thyroid calcitonin production in young mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28XjvFOrur8%3D&md5=3405dffe98997995ab17e5a43a475c18CAS | 16497798PubMed |

Linher, K., Wu, D., and Li, J. (2007). Glial cell line-derived neurotrophic factor: an intraovarian factor that enhances oocyte developmental competence in vitro. Endocrinology 148, 4292–4301.
Glial cell line-derived neurotrophic factor: an intraovarian factor that enhances oocyte developmental competence in vitro.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXpslCntb0%3D&md5=82b8f175d88c38a11c1c33f633968ecfCAS | 17540724PubMed |

Liu, Y., May, N. R., and Fan, C. M. (2001). Growth arrest specific gene 1 is a positive growth regulator for the cerebellum. Dev. Biol. 236, 30–45.
Growth arrest specific gene 1 is a positive growth regulator for the cerebellum.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXltFGlsLY%3D&md5=55edcc75c4c5b6ed0d11b60131b1717bCAS | 11456442PubMed |

Marques, G., and Fan, C. M. (2002). Growth arrest specific gene 1: a fuel for driving growth in the cerebellum. Cerebellum 1, 259–263.
Growth arrest specific gene 1: a fuel for driving growth in the cerebellum.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXjtFaquw%3D%3D&md5=a693080c1026667751bde11e86ff0c64CAS | 12879964PubMed |

Millán, J. L. (2013). The role of phosphatases in the initiation of skeletal mineralization. Calcif. Tissue Int. 93, 299–306.
The role of phosphatases in the initiation of skeletal mineralization.Crossref | GoogleScholarGoogle Scholar | 23183786PubMed |

Miwa, J. M., Stevens, T. R., King, S. L., Caldarone, B. J., Ibanez-Tallon, I., Xiao, C., Fitzsimonds, R. M., Pavlides, C., Lester, H. A., Picciotto, M. R., and Heintz, N. (2006). The prototoxin lynx1 acts on nicotinic acetylcholine receptors to balance neuronal activity and survival in vivo. Neuron 51, 587–600.
The prototoxin lynx1 acts on nicotinic acetylcholine receptors to balance neuronal activity and survival in vivo.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28XhtVWntL3F&md5=76c8923bdc22f4bb89862b62e186637cCAS | 16950157PubMed |

Miyado, K., Yamada, G., Yamada, S., Hasuwa, H., Nakamura, Y., Ryu, F., Suzuki, K., Kosai, K., Inoue, K., Ogura, A., Okabe, M., and Mekada, E. (2000). Requirement of CD9 on the egg plasma membrane for fertilization. Science 287, 321–324.
Requirement of CD9 on the egg plasma membrane for fertilization.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXlvVOmsw%3D%3D&md5=57d7bed29f6f5e84ce8954acfbb0457aCAS | 10634791PubMed |

Moore, M. L., Teitell, M. A., Kim, Y., Watabe, T., Reiter, R. E., Witte, O. N., and Dubey, P. (2008). Deletion of PSCA increases metastasis of TRAMP-induced prostate tumors without altering primary tumor formation. Prostate 68, 139–151.
Deletion of PSCA increases metastasis of TRAMP-induced prostate tumors without altering primary tumor formation.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXislCmsrc%3D&md5=029a73ad6c8a01941bd90ad863437079CAS | 18044730PubMed |

Niederkofler, V., Salie, R., Sigrist, M., and Arber, S. (2004). Repulsive guidance molecule (RGM) gene function is required for neural tube closure but not retinal topography in the mouse visual system. J. Neurosci. 24, 808–818.
Repulsive guidance molecule (RGM) gene function is required for neural tube closure but not retinal topography in the mouse visual system.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXhsFOisbk%3D&md5=97b3cbb4c7d39809b3d9836e98cbd24eCAS | 14749425PubMed |

Nishimura-Akiyoshi, S., Niimi, K., Nakashiba, T., and Itohara, S. (2007). Axonal netrin-Gs transneuronally determine lamina-specific subdendritic segments. Proc. Natl Acad. Sci. USA 104, 14 801–14 806.
Axonal netrin-Gs transneuronally determine lamina-specific subdendritic segments.Crossref | GoogleScholarGoogle Scholar |

Nosten-Bertrand, M., Errington, M. L., Murphy, K. P., Tokugawa, Y., Barboni, E., Kozlova, E., Michalovich, D., Morris, R. G., Silver, J., Stewart, C. L., Bliss, T. V., and Morris, R. J. (1996). Normal spatial learning despite regional inhibition of LTP in mice lacking Thy-1. Nature 379, 826–829.
Normal spatial learning despite regional inhibition of LTP in mice lacking Thy-1.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK28XhsVWksr4%3D&md5=5f9b921caaef6c1338289e3622e6edf9CAS | 8587606PubMed |

Oh, J., Takahashi, R., Kondo, S., Mizoguchi, A., Adachi, E., Sasahara, R. M., Nishimura, S., Imamura, Y., Kitayama, H., Alexander, D. B., Ide, C., Horan, T. P., Arakawa, T., Yoshida, H., Nishikawa, S., Itoh, Y., Seiki, M., Itohara, S., Takahashi, C., and Noda, M. (2001). The membrane-anchored MMP inhibitor RECK is a key regulator of extracellular matrix integrity and angiogenesis. Cell 107, 789–800.
The membrane-anchored MMP inhibitor RECK is a key regulator of extracellular matrix integrity and angiogenesis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38XhtFymuw%3D%3D&md5=8406e6f01ab44f02a496db196eb0c4b7CAS | 11747814PubMed |

Paratcha, G., and Ledda, F. (2008). GDNF and GFRalpha: a versatile molecular complex for developing neurons. Trends Neurosci. 31, 384–391.
GDNF and GFRalpha: a versatile molecular complex for developing neurons.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXptVentbs%3D&md5=fec0bf0a76a81c3fa0a35873d338006dCAS | 18597864PubMed |

Pasterkamp, R. J., Peschon, J. J., Spriggs, M. K., and Kolodkin, A. L. (2003). Semaphorin 7A promotes axon outgrowth through integrins and MAPKs. Nature 424, 398–405.
Semaphorin 7A promotes axon outgrowth through integrins and MAPKs.Crossref | GoogleScholarGoogle Scholar | 12879062PubMed |

Pitari, G., Malergue, F., Martin, F., Philippe, J. M., Massucci, M. T., Chabret, C., Maras, B., Dupre, S., Naquet, P., and Galland, F. (2000). Pantetheinase activity of membrane-bound Vanin-1: lack of free cysteamine in tissues of Vanin-1 deficient mice. FEBS Lett. 483, 149–154.
Pantetheinase activity of membrane-bound Vanin-1: lack of free cysteamine in tissues of Vanin-1 deficient mice.Crossref | GoogleScholarGoogle Scholar | 11042271PubMed |

Qin, X., Hu, W., Song, W., Grubissich, L., Hu, X., Wu, G., Ferris, S., Dobarro, M., and Halperin, J. A. (2009). Generation and phenotyping of mCd59a and mCd59b double-knockout mice. Am. J. Hematol. 84, 65–70.
Generation and phenotyping of mCd59a and mCd59b double-knockout mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1MXjt1Cgtr0%3D&md5=c40ca2a27e7e38657c713216cab09046CAS | 19051264PubMed |

Rikimaru, A., Komori, K., Sakamoto, T., Ichise, H., Yoshida, N., Yana, I., and Seiki, M. (2007). Establishment of an MT4-MMP-deficient mouse strain representing an efficient tracking system for MT4-MMP/MMP-17 expression in vivo using beta-galactosidase. Genes Cells 12, 1091–1100.
Establishment of an MT4-MMP-deficient mouse strain representing an efficient tracking system for MT4-MMP/MMP-17 expression in vivo using beta-galactosidase.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXhtVyktLbI&md5=fbfd7ffee310091a463f835255b72e31CAS | 17825051PubMed |

Rossi, J., Luukko, K., Poteryaev, D., Laurikainen, A., Sun, Y. F., Laakso, T., Eerikainen, S., Tuominen, R., Lakso, M., Rauvala, H., Arumae, U., Pasternack, M., Saarma, M., and Airaksinen, M. S. (1999). Retarded growth and deficits in the enteric and parasympathetic nervous system in mice lacking GFR alpha2, a functional neurturin receptor. Neuron 22, 243–252.
Retarded growth and deficits in the enteric and parasympathetic nervous system in mice lacking GFR alpha2, a functional neurturin receptor.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXhs12ksr8%3D&md5=10e2d66e18744a63c8aada0af1c46e1aCAS | 10069331PubMed |

Rothe, J., Lesslauer, W., Lotscher, H., Lang, Y., Koebel, P., Kontgen, F., Althage, A., Zinkernagel, R., Steinmetz, M., and Bluethmann, H. (1993). Mice lacking the tumour necrosis factor receptor 1 are resistant to TNF-mediated toxicity but highly susceptible to infection by Listeria monocytogenes. Nature 364, 798–802.
Mice lacking the tumour necrosis factor receptor 1 are resistant to TNF-mediated toxicity but highly susceptible to infection by Listeria monocytogenes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK3sXmtVWmurw%3D&md5=071af895e96eb9c1c4341e4fcb869385CAS | 8395024PubMed |

Russell, I. J., Legan, P. K., Lukashkina, V. A., Lukashkin, A. N., Goodyear, R. J., and Richardson, G. P. (2007). Sharpened cochlear tuning in a mouse with a genetically modified tectorial membrane. Nat. Neurosci. 10, 215–223.
Sharpened cochlear tuning in a mouse with a genetically modified tectorial membrane.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXptlWntA%3D%3D&md5=8c1541208d23524742d51e119f88aa9bCAS | 17220887PubMed |

Schneider, C., King, R. M., and Philipson, L. (1988). Genes specifically expressed at growth arrest of mammalian cells. Cell 54, 787–793.
Genes specifically expressed at growth arrest of mammalian cells.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaL1MXit1Khurc%3D&md5=45a29c04b6eba65488dae0d80e467f62CAS | 3409319PubMed |

Schueler-Furman, O., Glick, E., Segovia, J., and Linial, M. (2006). Is GAS1 a co-receptor for the GDNF family of ligands? Trends Pharmacol. Sci. 27, 72–77.
Is GAS1 a co-receptor for the GDNF family of ligands?Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28XhtV2lsLw%3D&md5=89e505eeb97633d5c4835f6c2a388e6fCAS | 16406089PubMed |

Scott, R. P., and Ibanez, C. F. (2001). Determinants of ligand binding specificity in the glial cell line-derived neurotrophic factor family receptor alpha S. J. Biol. Chem. 276, 1450–1458.
Determinants of ligand binding specificity in the glial cell line-derived neurotrophic factor family receptor alpha S.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXmtV2hsQ%3D%3D&md5=306162b4ce3e8b26a5a03e2ee4a93573CAS | 11018032PubMed |

Shah, G. N., Ulmasov, B., Waheed, A., Becker, T., Makani, S., Svichar, N., Chesler, M., and Sly, W. S. (2005). Carbonic anhydrase IV and XIV knockout mice: roles of the respective carbonic anhydrases in buffering the extracellular space in brain. Proc. Natl Acad. Sci. USA 102, 16 771–16 776.
Carbonic anhydrase IV and XIV knockout mice: roles of the respective carbonic anhydrases in buffering the extracellular space in brain.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXht1KgsL3O&md5=9f3e56974500d0cf020e38d8456d96e9CAS |

Sjöstrand, D., and Ibáñez, C. F. (2008). Insights into GFRalpha1 regulation of neural cell adhesion molecule (NCAM) function from structure–function analysis of the NCAM/GFRalpha1 receptor complex. J. Biol. Chem. 283, 13 792–13 798.
Insights into GFRalpha1 regulation of neural cell adhesion molecule (NCAM) function from structure–function analysis of the NCAM/GFRalpha1 receptor complex.Crossref | GoogleScholarGoogle Scholar |

Stebel, M., Vatta, P., Ruaro, M. E., Del Sal, G., Parton, R. G., and Schneider, C. (2000). The growth suppressing gas1 product is a GPI-linked protein. FEBS Lett. 481, 152–158.
The growth suppressing gas1 product is a GPI-linked protein.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXmtlGqsL8%3D&md5=a16ce5021b83e5564b212ab1a706b309CAS | 10996315PubMed |

Sun, X., Funk, C. D., Deng, C., Sahu, A., Lambris, J. D., and Song, W. C. (1999). Role of decay-accelerating factor in regulating complement activation on the erythrocyte surface as revealed by gene targeting. Proc. Natl Acad. Sci. USA 96, 628–633.
Role of decay-accelerating factor in regulating complement activation on the erythrocyte surface as revealed by gene targeting.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK1MXmtlCksA%3D%3D&md5=075e579814a4b39cfe6f24eca7c88b42CAS | 9892684PubMed |

Takeda, Y., Akasaka, K., Lee, S., Kobayashi, S., Kawano, H., Murayama, S., Takahashi, N., Hashimoto, K., Kano, M., Asano, M., Sudo, K., Iwakura, Y., and Watanabe, K. (2003). Impaired motor coordination in mice lacking neural recognition molecule NB-3 of the contactin/F3 subgroup. J. Neurobiol. 56, 252–265.
Impaired motor coordination in mice lacking neural recognition molecule NB-3 of the contactin/F3 subgroup.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXntVWgtb0%3D&md5=70299069e83c30a779778d2a3a85c9feCAS | 12884264PubMed |

Tomac, A. C., Grinberg, A., Huang, S. P., Nosrat, C., Wang, Y., Borlongan, C., Lin, S. Z., Chiang, Y. H., Olson, L., Westphal, H., and Hoffer, B. J. (1999). Glial cell line-derived neurotrophic factor receptor alpha1 availability regulates glial cell line-derived neurotrophic factor signaling: evidence from mice carrying one or two mutated alleles. Neuroscience 95, 1011–1023.
Glial cell line-derived neurotrophic factor receptor alpha1 availability regulates glial cell line-derived neurotrophic factor signaling: evidence from mice carrying one or two mutated alleles.Crossref | GoogleScholarGoogle Scholar |

Tansey, M. G., Baloh, R. H., Milbrandt, J., and Johnson, E. M. (2000). GFRalpha-mediated localization of RET to lipid rafts is required for effective downstream signaling, differentiation, and neuronal survival. Neuron 25, 611–623.
GFRalpha-mediated localization of RET to lipid rafts is required for effective downstream signaling, differentiation, and neuronal survival.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXisFKrtb4%3D&md5=7e7c37bff1f649de1dd4c7e7d89da790CAS | 10774729PubMed |

Turnberg, D., Botto, M., Warren, J., Morgan, B. P., Walport, M. J., and Cook, H. T. (2003). CD59a deficiency exacerbates accelerated nephrotoxic nephritis in mice. J. Am. Soc. Nephrol. 14, 2271–2279.
CD59a deficiency exacerbates accelerated nephrotoxic nephritis in mice.Crossref | GoogleScholarGoogle Scholar | 12937303PubMed |

Wang, Z. Q., Auer, B., Stingl, L., Berghammer, H., Haidacher, D., Schweiger, M., and Wagner, E. F. (1995). Mice lacking ADPRT and poly(ADP-ribosyl)ation develop normally but are susceptible to skin disease. Genes Dev. 9, 509–520.
Mice lacking ADPRT and poly(ADP-ribosyl)ation develop normally but are susceptible to skin disease.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK2MXksFOksLg%3D&md5=27758f549e95fa5a3133547df49a4eebCAS | 7698643PubMed |

Watanabe, H., and Kondoh, G. (2011). Mouse sperm undergo GPI-anchored protein release associated with lipid raft reorganization and acrosome reaction to acquire fertility. J. Cell Sci. 124, 2573–2581.
Mouse sperm undergo GPI-anchored protein release associated with lipid raft reorganization and acrosome reaction to acquire fertility.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXht1Skt7%2FO&md5=feabc4bd0606058bbe97958b6a3157b9CAS | 21750187PubMed |

Weinstock, P. H., Bisgaier, C. L., Aalto-Setala, K., Radner, H., Ramakrishnan, R., Levak-Frank, S., Essenburg, A. D., Zechner, R., and Breslow, J. L. (1995). Severe hypertriglyceridemia, reduced high density lipoprotein, and neonatal death in lipoprotein lipase knockout mice. Mild hypertriglyceridemia with impaired very low density lipoprotein clearance in heterozygotes. J. Clin. Invest. 96, 2555–2568.
Severe hypertriglyceridemia, reduced high density lipoprotein, and neonatal death in lipoprotein lipase knockout mice. Mild hypertriglyceridemia with impaired very low density lipoprotein clearance in heterozygotes.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DyaK28XmtVejsw%3D%3D&md5=d8e7adc1b6be8d73dc3c2ccd18356e8aCAS | 8675619PubMed |

Wlodarczyk, B., Spiegelstein, O., Gelineau-van Waes, J., Vorce, R. L., Lu, X., Le, C. X., and Finnell, R. H. (2001). Arsenic-induced congenital malformations in genetically susceptible folate binding protein-2 knockout mice. Toxicol. Appl. Pharmacol. 177, 238–246.
Arsenic-induced congenital malformations in genetically susceptible folate binding protein-2 knockout mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3MXpt12rtro%3D&md5=93091cbb047c1e510eba40ccf04fafdbCAS | 11749123PubMed |

Xia, Y., Sidis, Y., Mukherjee, A., Samad, T. A., Brenner, G., Woolf, C. J., Lin, H. Y., and Schneyer, A. (2005). Localization and action of Dragon (repulsive guidance molecule b), a novel bone morphogenetic protein coreceptor, throughout the reproductive axis. Endocrinology 146, 3614–3621.
Localization and action of Dragon (repulsive guidance molecule b), a novel bone morphogenetic protein coreceptor, throughout the reproductive axis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2MXntVahsr4%3D&md5=b258e947e5173e8a77f489e731b09b46CAS | 15890774PubMed |

Xia, Y., Cortez-Retamozo, V., Niederkofler, V., Salie, R., Chen, S., Samad, T. A., Hong, C. C., Arber, S., Vyas, J. M., Weissleder, R., Pittet, M. J., and Lin, H. Y. (2011). Dragon (repulsive guidance molecule b) inhibits IL-6 expression in macrophages. J. Immunol. 186, 1369–1376.
Dragon (repulsive guidance molecule b) inhibits IL-6 expression in macrophages.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BC3MXnsVKrtQ%3D%3D&md5=f75a33401ce62e7d222baa1f5fe3b606CAS | 21187450PubMed |

Yamaguchi, T., Hirota, K., Nagahama, K., Ohkawa, K., Takahashi, T., Nomura, T., and Sakaguchi, S. (2007). Control of immune responses by antigen-specific regulatory T cells expressing the folate receptor. Immunity 27, 145–159.
Control of immune responses by antigen-specific regulatory T cells expressing the folate receptor.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2sXosFeqt74%3D&md5=f09af4aee15b0d94711a7c8326b1dc75CAS | 17613255PubMed |

Yamamoto, K., Yoshida, K., Miyagoe, Y., Ishikawa, A., Hanaoka, K., Nomoto, S., Kaneko, K., Ikeda, S., and Takeda, S. (2002). Quantitative evaluation of expression of iron-metabolism genes in ceruloplasmin-deficient mice. Biochim. Biophys. Acta 1588, 195–202.
Quantitative evaluation of expression of iron-metabolism genes in ceruloplasmin-deficient mice.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD38XotVKqurY%3D&md5=b6d0c1874b95915721c750f0aba3d6edCAS | 12393173PubMed |

Yoshitake, H., Tsukamoto, H., Maruyama-Fukushima, M., Takamori, K., Ogawa, H., and Araki, Y. (2008). TEX101, a germ cell-marker glycoprotein, is associated with lymphocyte antigen 6 complex locus k within the mouse testis. Biochem. Biophys. Res. Commun. 372, 277–282.
TEX101, a germ cell-marker glycoprotein, is associated with lymphocyte antigen 6 complex locus k within the mouse testis.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD1cXmvFent7k%3D&md5=1dc1a4af72f3aa952f8d8c271ae75a16CAS | 18503752PubMed |

Yu, S., Michie, S. A., and Lowe, A. W. (2004). Absence of the major zymogen granule membrane protein, GP2, does not affect pancreatic morphology or secretion. J. Biol. Chem. 279, 50 274–50 279.
Absence of the major zymogen granule membrane protein, GP2, does not affect pancreatic morphology or secretion.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXhtVWjs7nK&md5=651d94b86efdfbead3cadaef2cee0b14CAS |

Zhao, X., Ueba, T., Christie, B. R., Barkho, B., McConnell, M. J., Nakashima, K., Lein, E. S., Eadie, B. D., Willhoite, A. R., Muotri, A. R., Summers, R. G., Chun, J., Lee, K. F., and Gage, F. H. (2003). Mice lacking methyl-CpG binding protein 1 have deficits in adult neurogenesis and hippocampal function. Proc. Natl Acad. Sci. USA 100, 6777–6782.
Mice lacking methyl-CpG binding protein 1 have deficits in adult neurogenesis and hippocampal function.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3sXktlygsbo%3D&md5=f61259348058efc923fe0cfd20301835CAS | 12748381PubMed |

Zhou, Z., Apte, S. S., Soininen, R., Cao, R., Baaklini, G. Y., Rauser, R. W., Wang, J., Cao, Y., and Tryggvason, K. (2000). Impaired endochondral ossification and angiogenesis in mice deficient in membrane-type matrix metalloproteinase I. Proc. Natl Acad. Sci. USA 97, 4052–4057.
Impaired endochondral ossification and angiogenesis in mice deficient in membrane-type matrix metalloproteinase I.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD3cXislSgtLo%3D&md5=c79747ee49f8e7fc32a9efec045d590aCAS | 10737763PubMed |

Zimmerman, U. J., Wang, P., Zhang, X., Bogdanovich, S., and Forster, R. (2004). Anti-oxidative response of carbonic anhydrase III in skeletal muscle. IUBMB Life 56, 343–347.
Anti-oxidative response of carbonic anhydrase III in skeletal muscle.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD2cXnslSmsrg%3D&md5=1cad4ef7ffd3eed5bb88979c44bf076bCAS | 15370882PubMed |

Ziyyat, A., Rubinstein, E., Monier-Gavelle, F., Barraud, V., Kulski, O., Prenant, M., Boucheix, C., Bomsel, M., and Wolf, J. P. (2006). CD9 controls the formation of clusters that contain tetraspanins and the integrin alpha 6 beta 1, which are involved in human and mouse gamete fusion. J. Cell Sci. 119, 416–424.
CD9 controls the formation of clusters that contain tetraspanins and the integrin alpha 6 beta 1, which are involved in human and mouse gamete fusion.Crossref | GoogleScholarGoogle Scholar | 1:CAS:528:DC%2BD28Xit1Klsbo%3D&md5=2a7c53132a14f8c52eb4ae18efae108aCAS | 16418227PubMed |