Uterine selection of human embryos at implantation

Jan J Brosens, Madhuri S Salker, Gijs Teklenburg, Jaya Nautiyal, Scarlett Salter, Emma S Lucas, Jennifer H Steel, Mark Christian, Yi-Wah Chan, Carolien M Boomsma, Jonathan D Moore, Geraldine M Hartshorne, Sandra Sućurović, Biserka Mulac-Jericevic, Cobi J Heijnen, Siobhan Quenby, Marian J Groot Koerkamp, Frank C P Holstege, Anatoly Shmygol, Nick S Macklon, Jan J Brosens, Madhuri S Salker, Gijs Teklenburg, Jaya Nautiyal, Scarlett Salter, Emma S Lucas, Jennifer H Steel, Mark Christian, Yi-Wah Chan, Carolien M Boomsma, Jonathan D Moore, Geraldine M Hartshorne, Sandra Sućurović, Biserka Mulac-Jericevic, Cobi J Heijnen, Siobhan Quenby, Marian J Groot Koerkamp, Frank C P Holstege, Anatoly Shmygol, Nick S Macklon

Abstract

Human embryos frequently harbor large-scale complex chromosomal errors that impede normal development. Affected embryos may fail to implant although many first breach the endometrial epithelium and embed in the decidualizing stroma before being rejected via mechanisms that are poorly understood. Here we show that developmentally impaired human embryos elicit an endoplasmic stress response in human decidual cells. A stress response was also evident upon in vivo exposure of mouse uteri to culture medium conditioned by low-quality human embryos. By contrast, signals emanating from developmentally competent embryos activated a focused gene network enriched in metabolic enzymes and implantation factors. We further show that trypsin, a serine protease released by pre-implantation embryos, elicits Ca(2+) signaling in endometrial epithelial cells. Competent human embryos triggered short-lived oscillatory Ca(2+) fluxes whereas low-quality embryos caused a heightened and prolonged Ca(2+) response. Thus, distinct positive and negative mechanisms contribute to active selection of human embryos at implantation.

Figures

Figure 1. Decidualizing endometrial cells are biosensors…
Figure 1. Decidualizing endometrial cells are biosensors of embryo quality.
(a) Venn diagram presenting the number of transcripts regulated in decidualizing HESCs significantly (P < 0.01) regulated in response to signals from developmentally competent embryos (DCE) and developmentally impaired embryos (DIE). (b) Gene Ontology classification of decidual genes regulated in response to soluble factors secreted by DIE. (c) Western blot analysis demonstrating the kinetics of HSC70 induction in primary HESC cultures decidualized with cAMP and MPA in a time-course lasting 8 d. β-ACTIN served as a loading control. A representative result from three different primary cultures is shown. Full length images are presented as Supplementary Information. (d) Primary HESCs were transfected with non-targeting (NT) siRNA or siRNA targeting HSC70, decidualized for 5 d, and then immunoblotted for HSC70. β-ACTIN served as a loading control. Full length images are presented as Supplementary Information. (e) HSC70 knockdown inhibits the secretion of decidual markers, PRL and IGFBP1, in primary HESC cultures differentiated in vitro for 5 d. The data represent mean (±SD) of triplicate experiments. * indicates P < 0.05, and *** P < 0.001. (f) The percentage of viable HESCs, transfected first with non-targeting (NT) siRNA or HSC70 targeting siRNA and then decidualized for 5 d, is presented relative to the number of viable cells in mock-transfected, undifferentiated cells (dotted line). The data represent the mean (±SD) of three biological repeat experiments.
Figure 2. HSC70 knockdown induces ER stress…
Figure 2. HSC70 knockdown induces ER stress in decidualizing stromal cells.
(a) Total cell lysates from primary HESC cultures, transfected first with non-targeting (NT) siRNA or siRNA targeting HSC70 and then decidualized with cAMP and MPA for 5 d, were immunoprobed for various proteins involved in UPR, ER stress, and autophagy as indicated. β-ACTIN served as a loading control. Full length images are presented as Supplementary Information. (b) HSC70 knockdown in HESCs promotes autophagosome formation. Primary cells cultured on chamber slides were transfected with either non-targeting (NT) or HSC70 targeting siRNA, decidualized for 5 d, stained for LC3B expression (green) and subjected to confocal microscopy. The nuclei were visualized with DAPI (blue). (c) Primary cultures were co-transfected with pcDNA3/XBP1-luc, pRL-sv40 and either siRNA targeting HSC70 or NT siRNA. The cells were left untreated or differentiated for 5 d before measuring luciferase activity. The results show the normalized mean firefly luciferase activity (±SD), expressed in relative light units (RLU), of four biological repeat experiments. *** indicates P < 0.001. (d) Confluent cultures were transfected as described in (c), left untreated (control) or decidualized for 5 d, and then exposed to 30 μl of unconditioned embryo culture medium (ECM) or media conditioned by DCEs or DIEs for 12 h. The results show normalized mean luciferase activity (±SD), expressed in relative light units (RLU), of three biological repeat experiments. *** indicates P < 0.001.
Figure 3. Competent pre-implantation human embryos actively…
Figure 3. Competent pre-implantation human embryos actively induce a supportive uterine environment.
(a) Venn diagram presenting the number of maternal genes significantly (P < 0.01) altered 24 h after exposure of mouse uterus to unconditioned embryo culture medium or medium conditioned by either developmentally competent or impaired human embryos (DCE and DIE, respectively). (b) Tryptic activity was measured in 16 cultures containing a total of 163 human embryos between day 4 to 6 of development. Activity in unconditioned medium was below dotted line. (c) Application of embryo-conditioned medium induces [Ca2+]i oscillations in Ishikawa cells. Black traces show [Ca2+]i recordings from individual cells in response to application of 1:20 diluted culture medium obtained from DCE (top panel) or DIE (bottom panel). Red traces represent average of all individual traces in each panel. (d) Total protein lysates obtained from Ishikawa cells 24 h after treatment with trypsin (10 nM) for 5 min were subjected to western blot analysis and immunoprobed for COX2. β-ACTIN served as a loading control. Full length images are presented as Supplementary Information.
Figure 4. Positive and negative mechanisms contribute…
Figure 4. Positive and negative mechanisms contribute to active selection of human embryos at implantation.
(A) Developmentally competent human embryos secrete evolutionary conserved serine proteases that activate epithelial Na+ channel (ENaC) expressed on luminal epithelial cells, triggering Ca2+ signalling and, ultimately, induction of genes involved in implantation and post-implantation embryo development. In concert, the decidualizing endometrium secretes serine protease inhibitors, such as murine SPINK3 and the human homolog SPINK1, to limit embryo-derived proteolytic activity. Note that acquisition of a secretory phenotype upon decidualization depends on massive expansion of the ER in HESCs. (B) By contrast, excessive protease activity emanating from developmentally compromised embryos that have breached the luminal epithelium down-regulates the expression of molecular chaperones in surrounding decidual cells, leading to accumulation of misfolded proteins and ER stress. This in turn compromises decidual cell functions and triggers tissue breakdown and early maternal rejection.

References

    1. Macklon N. S., Geraedts J. P. & Fauser B. C. Conception to ongoing pregnancy: the ‘black box’ of early pregnancy loss. Hum Reprod Update 8, 333–343 (2002).
    1. Rai R. & Regan L. Recurrent miscarriage. Lancet 368, 601–611 (2006).
    1. Fragouli E. et al. The origin and impact of embryonic aneuploidy. Hum Genet, 10.1007/s00439-013-1309-0 (2013).
    1. Mertzanidou A. et al. Microarray analysis reveals abnormal chromosomal complements in over 70% of 14 normally developing human embryos. Hum Reprod 28, 256–264, 10.1093/humrep/des362 (2013).
    1. Vanneste E. et al. Chromosome instability is common in human cleavage-stage embryos. Nat Med 15, 577–583 (2009).
    1. Quenby S., Vince G., Farquharson R. & Aplin J. Recurrent miscarriage: a defect in nature's quality control? Hum Reprod 17, 1959–1963 (2002).
    1. Rajcan-Separovic E. et al. Identification of copy number variants in miscarriages from couples with idiopathic recurrent pregnancy loss. Hum Reprod 25, 2913–2922, 10.1093/humrep/deq202 (2010).
    1. Stephenson M. D., Awartani K. A. & Robinson W. P. Cytogenetic analysis of miscarriages from couples with recurrent miscarriage: a case-control study. Hum Reprod 17, 446–451 (2002).
    1. Mansouri-Attia N. et al. Endometrium as an early sensor of in vitro embryo manipulation technologies. Proc Natl Acad Sci U S A 106, 5687–5692 (2009).
    1. Teklenburg G. et al. Natural selection of human embryos: decidualizing endometrial stromal cells serve as sensors of embryo quality upon implantation. PLoS One 5, e10258 (2010).
    1. Salker M. et al. Natural selection of human embryos: impaired decidualization of the endometrium disables embryo-maternal interactieons and causes recurrent pregnant loss. PLoS One 5, e10287, 10.1371/journal.pone.0010287 (2010).
    1. Salker M. S. et al. Deregulation of the serum- and glucocorticoid-inducible kinase SGK1 in the endometrium causes reproductive failure. Nat Med 17, 1509–1513, 10.1038/nm.2498 (2011).
    1. Salker M. S. et al. Disordered IL-33/ST2 activation in decidualizing stromal cells prolongs uterine receptivity in women with recurrent pregnancy loss. PLoS One 7, e52252, 10.1371/journal.pone.0052252 (2012).
    1. Weimar C. H. et al. Endometrial stromal cells of women with recurrent miscarriage fail to discriminate between high- and low-quality human embryos. PLoS One 7, e41424, 10.1371/journal.pone.0041424 (2012).
    1. Heijnen E. M. et al. A mild treatment strategy for in-vitro fertilisation: a randomised non-inferiority trial. Lancet 369, 743–749, 10.1016/S0140-6736(07)60360-2 (2007).
    1. Hartl F. U. & Hayer-Hartl M. Converging concepts of protein folding in vitro and in vivo. Nat Struct Mol Biol 16, 574–581, 10.1038/nsmb.1591 (2009).
    1. Kampinga H. H. & Craig E. A. The HSP70 chaperone machinery: J proteins as drivers of functional specificity. Nat Rev Mol Cell Biol 11, 579–592, 10.1038/nrm2941 (2010).
    1. Cloke B. et al. The androgen and progesterone receptors regulate distinct gene networks and cellular functions in decidualizing endometrium. Endocrinology 149, 4462–4474 (2008).
    1. Benbrook D. M. & Long A. Integration of autophagy, proteasomal degradation, unfolded protein response and apoptosis. Exp Oncol 34, 286–297 (2012).
    1. Kaushik S. & Cuervo A. M. Chaperone-mediated autophagy: a unique way to enter the lysosome world. Trends Cell Biol 22, 407–417, 10.1016/j.tcb.2012.05.006 (2012).
    1. Walter P. & Ron D. The unfolded protein response: from stress pathway to homeostatic regulation. Science 334, 1081–1086, 10.1126/science.1209038 (2011).
    1. Kabeya Y. et al. LC3, GABARAP and GATE16 localize to autophagosomal membrane depending on form-II formation. J Cell Sci 117, 2805–2812, 10.1242/jcs.01131 (2004).
    1. Iwawaki T., Akai R., Kohno K. & Miura M. A transgenic mouse model for monitoring endoplasmic reticulum stress. Nat Med 10, 98–102, 10.1038/nm970 (2004).
    1. Altmae S. et al. Research resource: interactome of human embryo implantation: identification of gene expression pathways, regulation, and integrated regulatory networks. Mol Endocrinol 26, 203–217, 10.1210/me.2011-1196 (2012).
    1. Han B. C., Xia H. F., Sun J., Yang Y. & Peng J. P. Retinoic acid-metabolizing enzyme cytochrome P450 26a1 (cyp26a1) is essential for implantation: functional study of its role in early pregnancy. J Cell Physiol 223, 471–479, 10.1002/jcp.22056 (2010).
    1. Ruan Y. C. et al. Activation of the epithelial Na+ channel triggers prostaglandin E(2) release and production required for embryo implantation. Nat Med 18, 1112–1117, 10.1038/nm.2771 (2012).
    1. Furuhashi M. & Hotamisligil G. S. Fatty acid-binding proteins: role in metabolic diseases and potential as drug targets. Nat Rev Drug Discov 7, 489–503, 10.1038/nrd2589 (2008).
    1. Yang H., Galea A., Sytnyk V. & Crossley M. Controlling the size of lipid droplets: lipid and protein factors. Current Opin Cell Biol 24, 509–516, 10.1016/j.ceb.2012.05.012 (2012).
    1. Kadowaki T. et al. Adiponectin and adiponectin receptors in insulin resistance, diabetes, and the metabolic syndrome. J Clin Invest 116, 1784–1792, 10.1172/JCI29126 (2006).
    1. Sharma N. et al. Implantation serine proteinase 1 exhibits mixed substrate specificity that silences signaling via proteinase-activated receptors. PLoS One 6, e27888, 10.1371/journal.pone.0027888 (2011).
    1. Sharma N. et al. Implantation Serine Proteinases heterodimerize and are critical in hatching and implantation. BMC Dev Biol 6, 61, 10.1186/1471-213X-6-61 (2006).
    1. Wong G. W., Yasuda S., Morokawa N., Li L. & Stevens R. L. Mouse chromosome 17A3.3 contains 13 genes that encode functional tryptic-like serine proteases with distinct tissue and cell expression patterns. J Biol Chem 279, 2438–2452, 10.1074/jbc.M308209200 (2004).
    1. Haig D. Genetic conflicts in human pregnancy. Q Rev Biol 68, 495–532 (1993).
    1. Chuong E. B., Rumi M. A., Soares M. J. & Baker J. C. Endogenous retroviruses function as species-specific enhancer elements in the placenta. Nat Genet 45, 325–329, 10.1038/ng.2553 (2013).
    1. Crespi B. & Semeniuk C. Parent-offspring conflict in the evolution of vertebrate reproductive mode. Am Nat 163, 635–653, 10.1086/382734 (2004).
    1. Emera D., Romero R. & Wagner G. The evolution of menstruation: a new model for genetic assimilation: explaining molecular origins of maternal responses to fetal invasiveness. BioEssays 34, 26–35, 10.1002/bies.201100099 (2012).
    1. Brosens J. J., Parker M. G., McIndoe A., Pijnenborg R. & Brosens I. A. A role for menstruation in preconditioning the uterus for successful pregnancy. Am J Obstet Gynecol 200, 615 e611–616 (2009).
    1. Emera D. et al. Convergent evolution of endometrial prolactin expression in primates, mice, and elephants through the independent recruitment of transposable elements. Mol Biol Evol 29, 239–247, 10.1093/molbev/msr189 (2012).
    1. Kim Y. J. et al. Transcriptional activation of Cidec by PPARgamma2 in adipocyte. Biochem Biophys Res Commun 377, 297–302, 10.1016/j.bbrc.2008.09.129 (2008).
    1. Arimura N., Horiba T., Imagawa M., Shimizu M. & Sato R. The peroxisome proliferator-activated receptor gamma regulates expression of the perilipin gene in adipocytes. J Biol Chem 279, 10070–10076, 10.1074/jbc.M308522200 (2004).
    1. Iwaki M. et al. Induction of adiponectin, a fat-derived antidiabetic and antiatherogenic factor, by nuclear receptors. Diabetes 52, 1655–1663 (2003).
    1. Forman B. M. et al. 15-Deoxy-delta 12, 14-prostaglandin J2 is a ligand for the adipocyte determination factor PPAR gamma. Cell 83, 803–812 (1995).
    1. Ohmuraya M. et al. Autophagic cell death of pancreatic acinar cells in serine protease inhibitor Kazal type 3-deficient mice. Gastroenterology 129, 696–705, 10.1016/j.gastro.2005.05.057 (2005).
    1. Kazal L. A., Spicer D. S. & Brahinsky R. A. Isolation of a crystalline trypsin inhibitor-anticoagulant protein from pancreas. J Am Chem Soc 70, 3034–3040 (1948).
    1. Brosens J. J., Hayashi N. & White J. O. Progesterone receptor regulates decidual prolactin expression in differentiating human endometrial stromal cells. Endocrinology 140, 4809–4820 (1999).
    1. Yang Y. H. et al. Normalization for cDNA microarray data: a robust composite method addressing single and multiple slide systematic variation. Nucleic Acids Res 30, e15 (2002).

Source: PubMed

3
订阅